genetic studies on recurrent miscarriage

94
GENETIC STUDIES ON RECURRENT MISCARRIAGE Milja Kaare Folkhälsan Institute of Genetics and Department of Medical Genetics University of Helsinki, Helsinki, Finland ACADEMIC DISSERTATION To be publicly discussed, with the permission of the Medical Faculty of the University of Helsinki, in Biomedicum Helsinki, Haartmaninkatu 8, lecture hall 2, on February 13 th 2009, at 1 pm. Helsinki 2009

Upload: others

Post on 06-Jan-2022

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Genetic Studies on Recurrent Miscarriage

GENETIC STUDIES ON RECURRENT MISCARRIAGE

Milja Kaare

Folkhälsan Institute of Genetics

and

Department of Medical Genetics

University of Helsinki,

Helsinki, Finland

ACADEMIC DISSERTATION

To be publicly discussed, with the permission of

the Medical Faculty of the University of Helsinki,

in Biomedicum Helsinki, Haartmaninkatu 8, lecture hall 2,

on February 13th

2009, at 1 pm.

Helsinki 2009

Page 2: Genetic Studies on Recurrent Miscarriage

Supervised by

Docent Kristiina Aittomäki, M.D., Ph.D.

Department of Clinical Genetics,

Helsinki University Central Hospital

Folkhälsan Institute of Genetics and

Department of Medical Genetics,

University of Helsinki

Reviewed by

Docent Leena Anttila, M.D., Ph.D.

Väestöliitto Fertility Clinics Ltd, Turku Clinic

Department of Obstetrics and Gynecology,

University of Turku

Docent Nina Horelli-Kuitunen, Ph.D., medical geneticist

Medix Laboratories Ltd, Espoo

Official opponent

Professor Seppo Heinonen, M.D., Ph.D.

Department of Obstetrics and Gynecology

University of Kuopio and

Kuopio University Hospital

ISBN 978-952-92-5027-1 (paperback)

ISBN 978-952-10-5259-0 (PDF)

http://ethesis.helsinki.fi

Yliopistopaino

Helsinki 2009

Page 3: Genetic Studies on Recurrent Miscarriage
Page 4: Genetic Studies on Recurrent Miscarriage

Contents

List of original publications _____________________________________________ 6

Abbreviations ________________________________________________________ 7

Abstract _____________________________________________________________ 9

Introduction ________________________________________________________ 11

Review of the literature ________________________________________________ 12

1 Early human embryonic development _______________________________ 12

1.1 Fertilization __________________________________________________ 12

1.2 Implantation __________________________________________________ 12

1.3 Placentation __________________________________________________ 14

1.4 Gastrulation and organogenesis ___________________________________ 14

2 Recurrent miscarriage ____________________________________________ 15

2.1 Identifiable causes of miscarriage _________________________________ 16

2.1.1 Risk factors for miscarriage __________________________________ 19

2.2 Genetic factors causing miscarriage _______________________________ 19

2.2.1 Chromosomal abnormalities __________________________________ 20

2.2.2 Skewed X chromosome inactivation ___________________________ 21

2.2.3 Y chromosome microdeletions ________________________________ 21

2.3.4 Gene mutations and polymorphisms ___________________________ 22

2.2.5 Aberrant placental gene expression ____________________________ 24

2.2.6 Epigenetic mechanisms _____________________________________ 25

3 Identification of genetic causes underlying RM ________________________ 26

3.1 Identification of disease genes ____________________________________ 27

3.1.1 The Human Genome Project _________________________________ 27

3.1.2 Traditional identification of disease genes _______________________ 28

3.1.3 The candidate gene approach _________________________________ 28

3.1.4 Mutation screening _________________________________________ 29

3.1.5 Confirming the nature of a new genetic variation _________________ 30

3.2 Mouse as a model organism _____________________________________ 30

3.3 Candidate genes for RM ________________________________________ 32

3.3.1 Amnionless _______________________________________________ 32

3.3.2 Thrombomodulin and Endothelial Protein C Receptor _____________ 35

3.3.3 p53 _____________________________________________________ 38

3.3.4 The mitochondrial genome ___________________________________ 39

Aims _______________________________________________________________ 43

Materials and methods ________________________________________________ 44

1 Study subjects ___________________________________________________ 44

1.1 Patients (I, II, III, IV, V) ______________________________________ 44

1.2 Controls (I, II, III, IV, V) ______________________________________ 44

Page 5: Genetic Studies on Recurrent Miscarriage

1.3 Placental samples (III, IV) _____________________________________ 45

1.4 Ethical issues (I, II, III, IV, V) __________________________________ 45

2 Methods ________________________________________________________ 45

2.1 Genealogic analysis (I) _______________________________________ 45

2.2 DNA extraction (I, III) ________________________________________ 45

2.3 Candidate gene analysis _______________________________________ 46

2.4 Immunohistochemical analysis of p53 (III) ________________________ 48

2.5 X chromosome inactivation analysis (V) __________________________ 48

2.6 Y chromosome microdeletion analysis (V) ________________________ 48

2.7 Statistical analysis (I, II, III, IV, V) ______________________________ 49

Results and discussion ________________________________________________ 50

1 Genealogic studies (I) _____________________________________________ 50

2 Candidate gene analysis (I, II, III, IV) _______________________________ 52

2.1 Amnionless (I) ________________________________________________ 52

2.1.1 Variations in the Amnionless gene _____________________________ 52

2.2 Thrombomodulin and Endothelial Protein C Receptor (II, unpublished) ___ 54

2.2.1 Variations in the TM gene ___________________________________ 54

2.2.2 Variations in the EPCR gene _________________________________ 55

2.2.3 TM/EPCR variations and outcome of heparin treatment ____________ 56

2.3 p53 (III, unpublished) __________________________________________ 57

2.3.1 Variations in the p53 gene ___________________________________ 57

2.3.2 p53 expression in RM placenta ________________________________ 58

2.4 Mitochondrial genes (IV) ________________________________________ 59

2.4.1 Variations in the mitochondrial genome _________________________ 60

2.4.2 Analysis of placental samples _________________________________ 62

3 The impact of polymorphisms on phenotype __________________________ 63

3.1 Variations in the nuclear genome (I, II, III, unpublished) _______________ 63

3.1.1 Possible splice site variations _________________________________ 65

3.1.2 Couple analysis _____________________________________________ 66

3.2 Variations in the mitochondrial genome (IV) ________________________ 67

4 Analysis of sex chromosome characteristics (V) _______________________ 68

4.1 X chromosome inactivation analysis _______________________________ 68

4.2 Y chromosome microdeletion analysis _____________________________ 71

5 Limitations of the study ___________________________________________ 73

Conclusions _________________________________________________________ 74

Acknowledgements ___________________________________________________ 77

References __________________________________________________________ 79

Page 6: Genetic Studies on Recurrent Miscarriage

6

List of original publications

This thesis is based on the following publications, which are referred to in the text by their

Roman numerals. In addition, some unpublished results are presented.

I Kaare M, Painter JN, Ulander VM, Kaaja R, Aittomaki K (2006) Variations

of the Amnionless gene in recurrent spontaneous abortions. Mol Hum

Reprod. 12:25-9.

II Kaare M*, Ulander VM*, Painter JN, Ahvenainen T, Kaaja R, Aittomaki K

(2007) Variations in the Thrombomodulin and Endothelial Protein C

Receptor genes in couples with recurrent spontaneous abortions. Hum

Reprod. 22:864-8.

III Kaare M, Bützow R, Ulander V-M, Kaaja R, Aittomäki K, Painter JN

(2009). Study of p53 gene mutations and placental expression in recurrent

miscarriage cases. RBM Online. E-pub ahead of print on 9 January.

IV Kaare M, Götz A, Ulander V-M, Ariansen S, Kaaja R, Suomalainen A,

Aittomäki K. Do mitochondrial mutations cause recurrent miscarriage?

Submitted.

V Kaare M, Painter JN, Ulander V-M, Kaaja R, Aittomäki K (2008) Sex

chromosome characteristics and recurrent miscarriage. Fertil Steril. 90:2328-

33.

* These two authors contributed equally to this work.

Publication II also appears in the thesis of Veli-Matti Ulander (2007)

The articles are reprinted with the permission of their copyright holders.

Page 7: Genetic Studies on Recurrent Miscarriage

7

Abbreviations

ACE angiotensin I-converting enzyme

AMN amnionless

APC activated protein C

APP amyloid beta precursor protein

AR androgen receptor

ATP adenosine triphosphate

AZF azoospermia factor

BAD BCL2 antagonist of cell death

BAX BCL2-associated X protein BID BH3-interacting domain death agonist

bp base pair

cDNA complementary DNA

CI confidence interval

CO2 Cytochrome c oxidase subunit II

cpm count per minute

CS carnegie stage

CTLA-4 cytotoxic T lymphocyte-associated 4

CUBN cubilin

CytB cytochrome b

dHPLC denaturing high performance liquid chromatography

DNA deoxyribonucleic acid

E embryonic day

EAA European Academy of Andrology

EMQN European Molecular Genetics Quality Network

EPCR endothelial protein C receptor

ESE exonic splicing enhancer

FasL Fas ligand

FSH-RO follicle stimulating hormone-resistant ovaries

HGP human genome project

HLA human leukocyte antigen

ICM inner cell mass

IFN-γ interferon-γ

IGS Imerslund-Gräsbäck syndrome

IL6/10 interleukin 6/10

ITGB3 integrin beta-3

MDM2 mouse double minute 2

MELAS mitochondrial encephalopathy with lactic acidosis and stroke-

like episodes

MERRF myoclonic epilepsy and ragged-red fibres

MLPA multiplex ligation-dependent probe amplification

MLS microphthalmia with linear skin defects syndrome

Page 8: Genetic Studies on Recurrent Miscarriage

8

mRNA messenger RNA

mtDNA mitochondrial DNA

MTHFR methylenetetrahydrofolate reductase

MUC1 mucin 1

ND2 NADH dehydrogenase subunit 2

OFD1 oral-facial-digital syndrome type 1

OR odds ratio

PAI-1 plasminogen activator inhibitor 1

PCR polymerase chain reaction

PGC primordial germ cells

PGD preimplantation genetic diagnosis

Q-PCR quantitative real-time PCR

PP14 placental protein 14

RM recurrent miscarriage

RNA ribonucleic acid

rRNA ribosomal RNA

SIFT sorting intolerant from tolerant

SNP single nucleotide polymorphism

SR-protein Ser/Arg-rich-protein

STS sequence tagged sites

Th-1 type 1 T-helper cell

TM thrombomodulin

tRNA transfer RNA

UTR untranslated region

VEGF vascular endothelial growth factor

VTE venous thromboembolism

XCI X chromosome inactivation

Page 9: Genetic Studies on Recurrent Miscarriage

9

Abstract

Recurrent miscarriage (RM) is defined as three or more consecutive pregnancy failures

and is estimated to affect ~1% of couples trying to conceive. The cause of RM remains

unknown in approximately 50% of cases although much work has been done in the past

to identify the underlying mechanisms. In this study, it was hypothesized that some of the

underlying factors yet to be discovered are genetic.

The aim of this study was to obtain new information on genetic causes of RM. The aim

was to search for mutations in genes known to cause miscarriage in animal models and

thereby find new genetic causes for unexplained miscarriages in humans. Four candidate

genes – Amnionless (AMN), Endothelial protein C receptor (EPCR), Thrombomodulin

(TM), and p53 – were chosen for the study as they have all been considered to be crucial

for normal embryonic development in the mouse. In addition, the mitochondrial genome

was studied because mitochondria are involved in processes important in early

development. Furthermore, sex chromosome characteristics suggested to underlie

miscarriage were also studied.

A total of 40 couples and 8 women with unexplained RM were collected for this study.

The patients were screened for mutations in the AMN, TM and EPCR genes using dHPLC.

Six interesting exonic and potential splice site disrupting variations were detected in these

genes, namely c.363G>A, c.843+11C>T, c.1339_1344dup, and c.1170-6C>T in AMN,

c.1728+23_+40del in TM, and c.323-9_336dup in EPCR. Two of these variations,

c.363G>A and c.843+11C>T were more frequent in RM women compared to controls.

However, their phenotypic effects cannot be determined without further investigations.

While most variations in the AMN, TM and EPCR genes were identified in both patients

and controls and are likely to be silent polymorphisms, others may have functional

significance or modifier effects in combination with mutations or polymorphisms in other

genes.

An association between the C11992A polymorphism of the p53 gene and RM was

detected. The results indicate that women carrying the C/A or A/A genotype have a

twofold higher risk for RM than women with a C/C genotype. This strengthens the results

of previous studies reporting that p53 sequence variations may cause miscarriage. The role

Page 10: Genetic Studies on Recurrent Miscarriage

10

of variation C11992A in embryonic development is, however, difficult to predict without

further studies. The role of p53 in RM was also investigated through expression analysis

of p53 in placental tissue. The results showed normal expression of p53 in the samples

studied.

When screening the mitochondrial genome a heteroplasmic mtDNA variation was

found in 13 (27%) RM women and 9 (19%) control women, which was unexpected as

heteroplasmic variations are reported to be rare. One novel variation within the coding

regions of the mitochondrial genome was detected. In addition, 18 previously reported

polymorphisms were identified. Although the detected variations are likely to be neutral

polymorphisms, a role in the aetiology of miscarriage cannot be excluded as some mtDNA

variations may be pathogenic only when a threshold is reached.

Recent publications have reported skewed X chromosome inactivation and Y

chromosome microdeletions to be associated with RM. Therefore, these sex chromosome

abnormalities in the context of RM were investigated. No association between skewed X

chromosome inactivation or Y chromosome microdeletions and RM in the Finnish

patients studied were detected.

In addition, data on ancestral birthplaces of the patients were collected to study any

possible geographic clustering, which would indicate a common predisposing factor. The

results showed clustering of the birthplaces in eastern Finland in a subset of patients. This

suggests a possibility of an enriched susceptibility gene which may contribute to RM.

These results encourage further studies to determine new genetic factors contributing to

the aetiology of RM.

Page 11: Genetic Studies on Recurrent Miscarriage

11

Introduction

Miscarriages are the most common complication of pregnancy, affecting approximately

15% of all clinically recognized pregnancies in the general population. The exact

frequency of miscarriages is, however, unknown as miscarriages frequently occur before

the woman is aware of her pregnancy. It is estimated that more pregnancies are lost

spontaneously than are actually carried to term (Rai and Regan, 2006; Stephenson and

Kutteh, 2007). Most of the miscarriages are sporadic and non-recurrent, and are often

caused by chromosome abnormalities in the foetus (Dhont, 2003). Recurrent miscarriage

(RM), defined as three consecutive pregnancy failures, is estimated to affect ~1% of all

couples trying to conceive (Li et al., 2002; Rai and Regan, 2006).

There are numerous factors that may cause RM, but the underlying problem often

remains undetected. Although much work has been done to identify the underlying

mechanisms, the cause of miscarriage can be identified in only ~50% of cases. The known

causes of RM include chromosomal and metabolic abnormalities, uterine anomalies, and

immunologic factors (Li et al., 2002; Rai and Regan, 2006). Even though RM is a

heterogeneous condition and the progress in identifying causative factors has been slow,

the repetitive pregnancy losses in some couples and the high percentage of unexplained

RM indicate that there are specific underlying causes yet to be found.

This study was conducted to gain further knowledge concerning genetic causes of RM.

Identification of the underlying factors is crucial for the development of more successful

treatment and improvement of the outcome of future pregnancies in women experiencing

RM. Genetic factors causing RM are, however, difficult to study because the foetus is lost

at an early stage of development and is therefore difficult to examine. Consequently, most

of the studies conducted on RM in general, and in this thesis work, are based on studying

the couples experiencing the miscarriages. We hypothesized that several genetic factors

such as mutations in genes, either identified from animal models or regulating functions

important in different stages of early foetal development, or abnormalities in the sex

chromosomes may underlie the miscarriages in the couples studied.

Page 12: Genetic Studies on Recurrent Miscarriage

12

Review of the literature

1 Early human embryonic development

1.1 Fertilization

The development of a human being starts from a single cell, a zygote, which results from

an ovum being fertilized by a spermatozoon. Fertilization occurs in the fallopian tube

within 48 hours of ovulation. After fertilization, a series of symmetrical cell divisions

create a mass of 12 to 16 totipotent cells, the morula, which is still enclosed within the

zona pellucida (Norwitz et al., 2001; Sariola et al., 2003). The morula enters the uterine

cavity approximately three days after fertilization. The appearance of a fluid filled inner

cavity within the mass of cells marks the transition from a morula to a blastocyst and is

accompanied by cellular differentiation (Norwitz et al., 2001; Wang and Dey, 2006). The

first differentiation event gives rise to trophoblasts, specialized epithelial cells that give

rise to extraembryonic structures, including the placenta. The remaining cells segregate at

one pole of the embryo to form the inner cell mass (ICM) (Figure 1). Within three days of

entering the uterine cavity, the embryo hatches from the zona pellucida, thereby exposing

its trophoblast cells, which enables implantation (Norwitz et al., 2001; Sariola et al., 2003;

Wang and Dey, 2006).

1.2 Implantation

The newly formed zygote has a discrete time frame in which it must prepare itself for

implantation. Preparation begins upon oocyte fertilization and takes approximately 6 to 7

days. During this time successive cell division occurs, and the outer trophoblast cells of

the blastocyst differentiate into cytotrophoblasts which diverge into villous and invasive

subtypes forming the building blocks of the placenta (Norwitz et al., 2001; Vitiello and

Patrizio, 2007). At the same time, when the blastocyst prepares itself for implantation, the

uterus becomes receptive. Uterine receptivity is defined as the state during the period of

endometrial maturation and proliferation when the blastocyst can become implanted.

Multiple signals synchronize the development of the blastocyst and the preparation of the

uterus. Of the many aspects of the synchronization process, the role of steroid hormones

is

Page 13: Genetic Studies on Recurrent Miscarriage

13

the best understood. Peptide hormones, growth factors, and cytokines also have roles in

the process. The cascade of signalling events that occur in both foetal and maternal tissues

establishes an appropriate environment critical to the development and survival of the

foetus (Norwitz et al., 2001; Wang and Dey, 2006; Vitiello and Patrizio, 2007).

Implantation occurs in three stages. The first stage is the initial adhesion of the

blastocyst to the uterine wall. The next stage is characterized

by increased physical

interaction between the blastocyst and the uterine epithelium. Shortly thereafter invasion

begins and syncytiotrophoblasts, the outer layer of the trophoblasts, penetrate the uterine

epithelium (Wang and Dey, 2006; Chavatte-Palmer and Guillomot, 2007; Vitiello and

Patrizio, 2007). By the 10th day after conception, the blastocyst is completely embedded

in the stromal tissue of the uterus as the uterine epithelium has regrown to cover the site of

implantation (Norwitz et al., 2001). The interaction between an activated blastocyst and a

receptive uterus is part of a complex process that leads to implantation

and the early stages

of placental development. Failure

to synchronize the processes involved in these

interactions results in failure of implantation (Cross et al., 1994; Wang and Dey, 2006;

Chavatte-Palmer and Guillomot, 2007).

Figure 1. The early stages of human development. A) An unfertilized egg cell is fertilized in the fallopian

tube by a sperm cell leading to B) a fertilized cell containing both the maternal and paternal pronuclei. The

cell undergoes cleavage to C) the 2-cell stage, D) the 4-cell stage and E) the 8-cell stage, which is reached

two to three days after fertilization. F) The morula enters the uterine cavity approximately three days after

fertilization. G) Implantation of the blastocyst, consisting of a blastocyst cavity, trophoblast cells and the

inner cell mass, into the uterine epithelium is initiated around seven days after fertilization. Modified from

Wang and Dey, 2006.

Page 14: Genetic Studies on Recurrent Miscarriage

14

1.3 Placentation

The placenta is the first organ to form during mammalian embryogenesis. The placenta is

a vital organ without which the embryo cannot survive in the uterine environment. The

function of the placenta during early gestation is primarily to mediate implantation of the

embryo into the uterus. After implantation, the major function of the placenta is to

mediate, as well as to regulate, nutrient uptake from the mother to the foetus. It also

establishes the interface for gas exchange between the maternal and foetal circulation.

Another important function of the placenta is the regulation of the maternal immune

response so that the foetal semi-allograft is tolerated during pregnancy. The placenta also

acts as an important source of hormones and growth factors that are needed for the initial

maternal recognition of pregnancy (Cross et al., 1994; Cross, 2006; Sadler et al., 2006).

Because the placenta is critical for survival it is very sensitive to disruption. Genetic or

environmental factors that affect the development of the placenta are associated with poor

pregnancy outcome. Abnormal expression of specific regulatory genes in the placenta and

abnormal functions of imprinted genes may cause placental dysfunction (Coan et al.,

2005). Abnormalities in placental functions can lead to a variety of problems including

implantation failure, placental insufficiency, foetal growth retardation and embryonic

death. Complications that become apparent relatively late in pregnancy may actually

reflect errors that occurred already during placental development (Cross et al., 1994;

Rossant and Cross, 2001; Cross, 2006).

1.4 Gastrulation and organogenesis

While the syncytiotrophoblasts start to penetrate into the wall of the uterus during

implantation, the embryoblast/ICM also continues developing. The embryoblast forms a

bilaminar embryo, composed of the epiblast and the hypoblast. The hypoblast forms the

yolk sac. The epiblast undergoes gastrulation at approximately day 16 following

fertilization. Gastrulation establishes the three germ layers—the endoderm, ectoderm, and

mesoderm—all of which will give rise to various organ systems. The mesoderm also

interacts with the trophoblast tissue to form the umbilical cord (Cross et al., 1994; Sadler

et al., 2006). At this point the actual embryonic phase starts and lasts from the third week

until the eighth week following fertilization. The organ systems differentiate at greatly

varying rates during this phase. For example, the circulatory system is largely functional at

Page 15: Genetic Studies on Recurrent Miscarriage

15

the end of this period, whereas the nervous system continues to undergo massive cell

division and is only beginning to establish functional connections (Sadler et al., 2006).

The remainder of human development, from weeks nine to thirty-eight, is called the foetal

period. During this period the embryo acquires its human appearance. The foetal period is

characterized by rapid growth and continued tissue and organ differentiation (Sariola et

al., 2003; Sadler et al., 2006).

2 Recurrent miscarriage

Human reproduction entails a fundamental paradox: although it is critical for the survival

of the species, the process is relatively inefficient. Maximal fecundity (the probability of

fertilization during one menstrual cycle) is approximately 30%. In addition, although much

research and many advances have been made in reproductive medicine during the past

decades, miscarriage, the spontaneous loss of a pregnancy before the foetus reaches

viability, remains the most common complication of pregnancy (Clark et al., 2001; Pandey

et al., 2004).

There are two types of miscarriage; sporadic and recurrent. In the general population

~15% of all clinically recognized pregnancies end in miscarriage (Stray-Pederson and

Stray-Perderson, 1984; Nybo Andersen et al., 2000). However, the rate of miscarriage

may be as high as 50-60% if the undetected, very early pregnancy losses occurring within

the first weeks following gestation are taken into account (Roberts and Lowe, 1975;

Wilcox et al., 1988). Of the pregnancies that are lost, ~70-75% are caused by failures of

implantation or early placentation and are therefore not clinically

recognized as

pregnancies. As a consequence, only about half of all conceptions advance into the second

trimester of pregnancy (Figure 2) (Rai and Regan, 2006).

RM is estimated to occur in 1-2% of all couples (Tulppala et al., 1993; Katz and

Kuller, 1994), but the exact prevalence of RM is dependent on its definition. RM is often

defined as the occurrence of three or more consecutive, clinically detectable pregnancy

failures before the 20th

week of gestation (Li et al., 2002b; Quenby et al., 2002; Pandey et

al., 2004). However, some studies have also included patients with two miscarriages

(Laskin et al., 1997; Stephenson et al., 1998) and in some studies the miscarriages were

Page 16: Genetic Studies on Recurrent Miscarriage

16

not necessarily consecutive. This increases the prevalence of RM up to 5% of all couples

trying to conceive (Rai and Regan, 2006).

Figure 2. Number of conceptions expected to reach different stages of pregnancy. Of all fertilized oocytes

only about half result in a livebirth. Most of the pregnancies are lost during early stages of development.

Modified from Rai and Regan, 2006.

2.1 Identifiable causes of miscarriage

The repetitive miscarriages in some couples suggest that certain women are at particular

risk of losing their pregnancy and that there must be an underlying explanation for this.

Even though much work has been done to identify these underlying mechanisms, the

aetiology of miscarriage is still unknown in many cases. Additionally, on an individual

level the exact reason for a particular miscarriage is rarely defined. Despite the extensive

medical testing and experimental treatments that many patients undergo, the cause often

remains unclear (Plouffe et al., 1992; Clifford et al., 1994; Carrington et al., 2005). The

main identifiable causes of miscarriage include the following categories (Table 1):

Uterine pathology; Uterine defects, such as uterine anomalies and fibroids, can

predispose to miscarriage by affecting implantation (Bulletti et al., 1996; Bajekal and Li,

2000).

Endocrine abnormalities; Following implantation, the maintenance of the pregnancy is

dependent on a range of endocrinological events that will eventually aid in the successful

growth and development of the endometrium and the foetus (Potdar and Konje, 2005;

Arredondo and Noble, 2006). Normal endometrial development and maturation is

dependent on steroid hormones, particularly oestrogen and progesterone. Therefore,

Conseption

Implantation

Recogniced

pregnancy

Livebirth

100

80

60

40

20

%

Page 17: Genetic Studies on Recurrent Miscarriage

17

miscarriage can be a result of subnormal hormone production or abnormal endometrial

response to circulating steroid hormones (Li et al., 2002a; Arredondo and Noble, 2006).

In addition, several other maternal endocrinological abnormalities such as uncontrolled

diabetes (Mills et al., 1988), high androgen levels (Okon et al., 1998; Bussen et al., 1999),

hyperprolactinaemia (Tal et al., 1991), thyroid dysfunction (Roberts and Murphy, 2000),

and obesity (Wang JX et al., 2002) have been implicated as aetiological factors for RM.

Immunological factors; Since the conceptus displays paternal gene products and

antigens, it is possible that the maternal immune system recognizes these as foreign,

resulting in an immune response (Dalton et al., 1998; Hill and Choi, 2000). Survival of the

semiallogenic foetus is dependent on suppression of the maternal immune response and

miscarriage may be a consequence of a failure of this suppression. It is assumed that both

maternal and embryonic regulating factors protect the embryo against an adverse maternal

immunological reaction. Maternal adaptation to the implanting embryo is needed for

successful establishment of the foetal-placental unit, and the embryo must be able to

stimulate an immunological reaction from the mother in order to become protected from

Table 1. Causes of RM and their estimated frequencies in women with RM (Stephenson and Kutteh, 2007;

Warren and Silver, 2008).

Factor Frequency

Uterine defects ~15%

Anatomic deformation

Uterine fibroids

Cervical incompetence

Endocrinologic ~10%

Abnormal hormone production

Diabetes

Thyroid dysfunction

Obesity

Immunologic ~15%

Aniphospholipid syndrome

Maternal autoimmune factors (acquired or inherited)

Thrombophilic ~10%

Prothrombotic state (acquired or inherited)

Environmental ~5%

Infections

Teratogens (alcohol, drugs, tobacco)

Parental chromosome ~5%

abnormality

Balanced translocations

Page 18: Genetic Studies on Recurrent Miscarriage

18

cytotoxic molecules (Bulletti et al., 1996; Li et al., 2002b; Pandey et al., 2004). The

mechanism by which the foetus escapes rejection by the maternal immune system is

unknown. However, differences in the concentration of circulating immunocompetent

cells, such as natural killer (NK) cells, and cytokines have been reported between RM

women and controls (Wegmann et al., 1993; Reinhard et al., 1998; Ntrivalas et al., 2001;

Yamada et al., 2001).

Prothrombic state; A prothrombotic state can be broadly defined as any condition,

whether inherited or acquired, that predisposes to venous or arterial thrombosis (Hiatt and

Lentz et al., 2002). It is postulated that maternal thrombophilia is a risk factor for

miscarriage because a successful pregnancy is dependent on satisfactory placental

development and sustained placental function is based on placental circulation (Kujovich,

2004). These processes require the establishment of an adequate foeto-maternal

circulatory system, which in turn may be disturbed by a prothrombotic state. Factor V

Leiden, mutations in the prothrombin gene, and protein C and protein S deficiencies are

hereditary prothrombotic disorders of clinical importance and associations between

increased risk of miscarriage and these thrombophilic disorders have been reported

(Preston et al., 1996; Rey et al., 2003; Kujovich, 2004).

One of the major acquired prothrombotic disorders is antiphospholipid antibody

syndrome, in which the body recognizes its own phospholipids as foreign and produces

antibodies against them. Lupus anticoagulant and anticardiolipin antibody are the two

known antiphospholipid antibodies that are associated with RM (Levine et al., 2002). It is

unknown exactly how the antiphospholipid antibody syndrome adversely affects

pregnancy, but one theory is that it may cause blood clots in the blood vessels of the

placenta, impairing placental function (Rand et al., 1997). However, antiphospholipid

syndrome is a treatable cause of RM. Various treatments, such as heparin, aspirin, and

intravenous immunoglobulin, have been used in attempts to improve the pregnancy

outcome of women with antiphospholipid syndrome (Empson et al., 2002; Farquharson et

al., 2002). Heparin is a thromboprophylactic agent that can bind to antiphospholipid

antibodies, thereby protecting the trophoblast and maternal vascular endothelium from

damage in early pregnancy. Later in pregnancy, when the intervillous circulation has been

established, heparin helps to decrease the risk of placental thrombosis and infarction

(Girardi, 2005, Bose et al., 2005).

Page 19: Genetic Studies on Recurrent Miscarriage

19

Environmental factors; Heavy metals, organic solvents, and ionizing radiation are

confirmed teratogens, and exposure to these can contribute to pregnancy loss (Dhont,

2003). Alcohol and cocaine are also confirmed teratogens, while caffeine and smoking are

suspected teratogens but their teratogenic impact is still controversial (Dominguez-Rojas

et al., 1994; Parazzini et al., 1998; Cnattingius et al., 2000; Kesmodel et al., 2002).

2.1.1 Risk factors for miscarriage

In addition to identifiable causes of miscarriage, other factors increasing the risk for

miscarriage have been reported. Increased maternal age is reported to be an independent

risk factor for miscarriage (Risch et al., 1988; Abdalla et al., 1993; Nybo Andersen et al.,

2000). The risk of mscarriage increases rapidly after the age of 35. The risk of a

miscarriage is 9% in women aged 20-24 years, while in women aged over 45 years the

risk is as high as 75% (Nybo Andersen et al., 2000). This is explained partly, but not

totally, by the association of maternal age with increased likelihood of chromosomal

abnormalities in the foetus (Munne et al., 1995; Snijders et al., 1999).

The outcome of previous pregnancies, especially of the first pregnancy, is also an

independent predictor of future pregnancy outcome. For young women who have never

experienced pregnancy loss the rate of miscarriage is as low as 5% (Regan et al., 1989).

The risk increases to approximately 30% for women with three or more losses but with a

previous live-born infant, and up to 50% for women with no live-born infants (Poland et

al., 1977), indicating that the risk of miscarriage increases with the number of previous

miscarriages. Additionally, a normal karyotype in a pregnancy loss is a predictor of

subsequent miscarriages because an abnormal foetal karyotype is associated with sporadic

foetal loss. The frequency of abnormal foetal karyotypes significantly decreases with the

number of miscarriages (Ogasawara et al., 2000).

2.2 Genetic factors causing miscarriage

In addition to the previously described identifiable causes of miscarriage, genetic factors

are a major cause of clinically recognized miscarriages. In the following sections, the

currently reported genetic causes associated with miscarriages are described.

Page 20: Genetic Studies on Recurrent Miscarriage

20

2.2.1 Chromosomal abnormalities

Most unbalanced chromosome aberrations result in severe phenotypes already in early

pregnancy and therefore lead to miscarriage. Consequently the incidence of foetal

unbalanced chromosomal abnormalities gradually decreases with duration of pregnancy to

less than 1% among live-born children (Stern et al., 1996; Stephenson et al., 2002).

Overall, 50-70% of all sporadic miscarriages are associated with cytogenetic abnormalities

in the foetus. The most frequent of these abnormalities is trisomy, followed by polyploidy,

monosomy X, and unbalanced translocation (Kalousek et al., 1993; Stephenson et al.,

2002). The most common trisomy is trisomy of chromosome 16, followed by trisomy of

chromosomes 21 and 22 (Stephenson and Kutteh, 2007). These chromosomal

abnormalities may be inherited or may arise de novo in the germ cells due to aberrant

oocyte spindle formation and meiotic division (Devine et al., 2000; Kuo, 2002).

In women with RM the rate of chromosomal abnormalities in the foetus is lower than

in women with sporadic miscarriage. The miscarriages in women with RM may be due to

de novo numerical chromosome abnormalities, in particular autosomal trisomies of

chromosomes 13, 14, 15, 16, 21, and 22, and monosomy X (Pandey et al., 2005; Warren

and Silver, 2008). In women with RM, however, the miscarriage may also be due to

recurrent chromosomal anomalies in the foetus resulting from a balanced aberration in one

of the parents inherited by the offspring in an unbalanced form. Cytogenetic screening of

couples with RM has revealed that parental chromosomal abnormalities occur in either

partner in 5-7% of couples with RM, while the rate in the normal population is

approximately 0.2% (Fryns and Buggenhout, 1998). Balanced translocations in either

parent, including reciprocal and Robertsonian translocations, are the most common

abnormalities, detected in about 4% of couples with RM. The other chromosomal

abnormalities include inversions and sex chromosome abnormalities (Tulppala et al.,

1993; Clifford et al., 1994; Fryns and Buggenhout, 1998). Even though the theoretical risk

for abnormal offspring for phenotypically normal carriers of balanced translocations is

approximately 50%, preimplantation genetic diagnosis (PGD) analysis has shown that the

risk varies greatly. For carriers of a balanced translocation the risk of conceiving a

chromosomally abnormal embryo, due to abnormal segregation at meiosis, varies from 20-

80%. The reproductive risk conferred by chromosome rearrangements is dependent on the

size, location, and type of rearrangement and whether it is carried by the woman or her

male partner (Munne et al., 2000; Otani et al., 2006; Lim C et al., 2008).

Page 21: Genetic Studies on Recurrent Miscarriage

21

2.2.2 Skewed X chromosome inactivation

X chromosome inactivation (XCI) is defined as the process where one of the two X

chromosomes in each somatic cell of a healthy female is inactivated during early

embryogenesis. The process is important to achieve appropriate dosage compensation for

X chromosomal genes (Lyon, 1961; Brown and Robinson, 2000). Females are therefore

mosaics for two cell types; cells with the paternal X chromosome as the active

X and cells

with the maternal X chromosome as the active X. Normally the process occurs randomly

on both the maternal and paternal X chromosome, resulting in a 50:50 distribution of the

two cell types. A significant deviation from this distribution is called skewed XCI (Puck

and Willard, 1998; Minks et al., 2008). Recently, several studies have shown that women

experiencing RM exhibit nonrandom XCI more often than control women (Lanasa et al.,

1999; Uehara et al., 2001; Kuo et al., 2008). Kuo et al. (2008) reported extremely skewed

XCI in 7.7% of RM women and in 1.3% of control women (p=0.01).

Skewed XCI, defined as preferential inactivation of one of the two X chromosomes,

may be the result of a chance event, or due to genetic factors involved in the X inactivation

process (Brown and Robinson, 2000). In some cases a non-random inactivation pattern

may be an indicator of an abnormality in the inactivated X, e.g. X chromosome

aberrations such as microdeletions or mutations in X chromosomal genes (Belmont, 1996;

Uehara et al., 2001). Deviation from random inactivation rates can also lead to the

expression of recessive X-linked traits in heterozygous females, who are expected to be

healthy carriers of the disease gene (Azofeifa et al., 1995). It has been proposed that X-

linked mutations which are lethal in males cause skewed XCI in female carriers. For

example, microphthalmia with linear skin defects syndrome (MLS) and oral-facial-digital

syndrome type 1 (OFD1) are previously described X-linked dominant male-lethal diseases

(Franco and Ballabio, 2006). Such mutations may result in loss of male conceptuses and

increased frequency of miscarriage in carrier females (Lanasa et al., 1999; Uehara et al.,

2001).

2.2.3 Y chromosome microdeletions

The majority of the testing for RM assesses the woman. The male partners, however, have

been less thoroughly investigated. Semen analysis is not generally included in the initial

assessment for RM. Yet, sperm integrity is required for sperm–egg interactions,

Page 22: Genetic Studies on Recurrent Miscarriage

22

fertilization, and early embryonic development and sperm quality is associated with the

embryo’s ability to reach the blastocyst stage and progress to implantation (Simerly et al.,

1995; Van Blerkom, 1996; Moomjy et al., 1999).

Microdeletions of the long arm of the human Y chromosome are associated with

variable spermatogenic failure (Vogt et al., 1996; Hopps et al., 2003). These

microdeletions occur in regions called AZFa, AZFb, and AZFc (more recently named

AZFa, P5/Proximal-P1, P5/Distal-P1, P4/Distal-P1, b2/b4-AZFc), containing a number of

genes (Hopps et al., 2003; Ferlin et al., 2007). Although ten recurrent deletions in the AZF

regions have been described in detail (Ferlin et al., 2007), surprisingly little is known

about the function of the individual genes and transcription units located within these

regions (Hopps et al., 2003; Foresta et al., 2005). Y chromosome microdeletions are found

in approximately 7% of men with very low sperm counts (Ferlin et al., 2007). These men

also often show a higher incidence of sperm chromosomal abnormalities (Foresta et al.,

2005). An elevated rate of sperm chromosome abnormalities has also been found in some

men of RM couples (Rubio et al., 1999; Egozcue et al., 2000). A recent study reports that

the incidence of sperm chromosome abnormalities is 16.5% in male partners of women

with RM, while the incidence is under 5% in controls (Al-Hassan et al., 2005).

Interestingly, a pilot study conducted to determine if Y chromosome microdeletions are

associated with RM, reported that a significant portion (82%) of male partners of women

with RM had Y chromosomal microdeletions, which may underlie the miscarriages

experienced by their partners (Dewan et al., 2006).

2.3.4 Gene mutations and polymorphisms

Mutations and polymorphisms in several genes are suggested to contribute to RM. Many

of these sequence variations are located in genes involved in immunological mechanisms

or blood coagulation pathways, but also genes involved in other functions have been

associated with RM.

Since the conceptus displays paternal gene products and antigens, it is possible that the

maternal immune system recognizes these as foreign, resulting in an immune response

(Dalton et al., 1998; Hill and Choi, 2000). In recent years, several genes involved in the

immune response have been associated with pregnancy loss. The CTL-A gene is a

Page 23: Genetic Studies on Recurrent Miscarriage

23

regulator of T-cell activation and an A/G polymorphism in exon 1 of the gene has been

reported to be associated with RM in the Chinese population. The frequency of the GG

genotype in RM patients was significantly higher compared with controls (48.8% versus

33.3%, p = 0.011) (Wang et al., 2005). Polymorphisms in the major histocompatibility

complex have also been associated with RM. A polymorphism in the HLA-G promoter

region, the HLA-G *0104 and *0105N alleles (Aldrich et al., 2001; Ober et al., 2003) and

the HLA-DRB1 *1505 allele have been associated with the immunopathogenesis of RM

(Takakuwa et al., 2003). In addition, a number of cytokine gene polymorphisms are

associated with pregnancy loss. Cytokines are critical immunoregulatory molecules which

play a crucial role in early embryo development and placentation. Associations have been

shown between IL-6, IL-10, and IFN-γ gene polymorphisms and RM (Daher et al., 2003;

Costeas et al., 2004; Prigoshin et al., 2004; von Linsingen et al., 2005).

It is assumed that maternal thrombophilia is a risk factor for miscarriage because a

successful pregnancy is dependent on an adequate foeto-maternal circulatory system,

satisfactory placental development and sustained placental function. These functions may

be disturbed by a prothrombotic state. Factor V and prothrombin gene mutations, protein

C and protein S deficiencies, cause hereditary prothrombotic disorders (Preston et al.,

1996; Kujovich, 2004). Associations between these thrombophilic disorders and an

increased risk of miscarriage have been reported. Depending on the study the Factor V

Leiden mutation and the prothrombin G20210A mutation are reported to increase the risk

for RM between 2-fold and 5.5-fold, and between 2.5-fold and 4.9-fold, respectively

(Souza et al., 1999; Foka et al., 2000; Reznikoff-Etiévan et al., 2001; Rey et al., 2003).

Protein C and protein S deficiencies are much rarer and they have been reported to

increase the risk of early RM ~3.5-fold and ~14.5-fold, respectively (Rey et al., 2003;

Kujovich, 2004). The MTHFR gene has an important role in homocysteine metabolism

and missense mutations C677T and A1298C are known to cause hyperhomocysteinemia.

Hyperhomocysteinemia is a risk factor for vascular thrombophilia and has also been

shown to associate with RM (Couto et al., 2005; Mtiraoui et al., 2006). In addition,

polymorphisms in the ITGB3 gene, which encodes a human platelet antigen, may increase

the risk of miscarriage by causing thrombophilia. The PLA2 polymorphism in the ITGB3

gene was shown to increase the risk of foetal loss 3.6-fold (Ruzzi et al., 2005).

In addition, various other genes have been associated with pregnancy loss. For

example, mutations in the SCO2 gene, needed for cytochrome c oxidase assembly, were

Page 24: Genetic Studies on Recurrent Miscarriage

24

detected in a compound heterozygous state in aborted foetuses and are suggested to cause

miscarriage (Tay et al., 2004). A novel A/G polymorphism in intron 6 of the eNOS gene,

which is active during implantation, was recently reported to be associated with RM

(Suryanarayana et al., 2006). A polymorphism in the VEGF gene, a well characterized

regulator of angiogenesis, was also recently reported as a new susceptibility factor for

RM. Homozygosity of the VEGF –1154A variation was more frequent in RM women

compared to controls and increased the risk of RM 2.7-fold (Coulam and Jeyendran,

2008).

2.2.5 Aberrant placental gene expression

The growth and differentiation of placental structures are regulated by many genes and the

placenta is very sensitive to genetic disruption, as reflected by the increasing list of

targeted mouse mutations in genes such as Mash2 and Hand1 that cause placental defects

and embryonic lethality (Rossant and Cross, 2001). Genetic abnormalities that affect the

development of the placenta can cause placental insufficiency, foetal growth retardation

and embryonic death, both in mice and humans (Rossant and Cross, 2001; Cross, 2006).

It is assumed that normal placentation and normal placental functions are mediated by

differential

gene expression in the placenta. It has also been suggested that genes

expressed in the chorionic villi act as critical regulators of placentation and hence of early

human development (Dizon-Townson et al., 2000; Baek, 2004). Recently, over 30 genes

that are expressed aberrantly in pregnancy failure have been identified. The genes showing

different levels of expression between controls and patients with RM can be mainly

grouped as immunity related, angiogenesis related, and apoptosis related genes (Baek et

al., 2002; Choi et al., 2003). Higher expression levels of apoptosis related genes such as

caspase 3, 6, 7, 8, 9, 10, 12, BAD, BAX, BID, Fas and FasL were shown

in chorionic villi

from patients with RM compared to controls (Choi et al, 2003). Immunity related genes

with aberrant gene expression in chorionic villi include PP14, MUC1 (Baek et al., 2002),

CD95 (Hoshimoto et al., 2002), and Annexin II (Aarli et al., 1997). These genes may be

involved in immunological foeto-maternal reactions. Fibronectin, Integrin and VEGF are

examples of angiogenesis related genes showing aberrant gene expression in RM placenta

(Choi et al, 2003). These genes are involved in the formation of blood vessels that are

crucial for placental and embryonic development.

Page 25: Genetic Studies on Recurrent Miscarriage

25

2.2.6 Epigenetic mechanisms

The term epigenetics refers to processes that regulate gene activity without affecting

the

DNA sequence and are heritable through cell division. Epigenetic modifications of DNA

such as methylation are important for genome function during development, and

disturbances in the highly coordinated epigenetic processes contribute to developmental

failures (Lucifero et al., 2004; Santos and Dean, 2004). In most mammalian species,

including humans, there are at least two developmental periods during which methylation

patterns are reprogrammed genome wide; germ cell development and preimplantation.

After genome wide demethylation of the genome in developing germ cells, remethylation,

which resets the parent-of-origin specific marks, occurs. The second phase of

reprogramming of methylation occurs during the preimplantation stage. The paternal

genome is actively demethylated within a few hours of fertilization in the human zygote,

whereas the maternal genome is passively demethylated by a replication dependent

mechanism after the two cell embryo stage. After the demethylation stage, de novo

methylation of the genome occurs in the developing embryo. This reprogramming is likely

to have a crucial role in establishing nuclear totipotency required for normal development

(Reik et al., 2001; Reik and Dean, 2001; Haaf, 2006).

Genomic imprinting is a mechanism which results in differential expression of

maternal and paternal genes. Only one of the inherited gene copies is active in an embryo

depending on the parental origin of the gene. DNA methylation in germ cells is critical for

imprinting as most imprinted genes have differentially methylated regions (Reik and

Dean, 2001; Swales and Spears, 2005). Anomalous expression of imprinted genes during

development is sometimes caused by imbalanced representation of maternal and paternal

genetic contributions, uniparental disomy, and is associated with developmental delay or

gestational loss (Mutter, 1997).

It is estimated that there are several hundred imprinted genes in the human genome. It

has been argued that imprinted genes play essential roles in controlling

the placental

supply of maternal nutrients to the foetus, by regulating the growth of the placenta and/or

the activity of

transplacental transport systems. In some cases,

errors in genomic

imprinting are embryo lethal while in others they lead to developmental disorders and

diseases (Reik and Dean, 2001; Lucifero et al., 2004; Swales and Spears, 2005).

Page 26: Genetic Studies on Recurrent Miscarriage

26

3 Identification of genetic causes underlying RM

Despite numerous investigation the cause of miscarriage remains unknown in

approximately 50% of cases (Plouffe et al., 1992; Tulppala et., al 1993, Clifford et al.,

1994). A number of possible aetiologies have been proposed to explain the RM in these

cases. The hypotheses, however, change over time and new findings have enriched our

understanding of possible mechanisms underlying RM. For example, specific HLA-G

alleles (Aldrich et al., 2001) and skewed X-inactivation (Lanasa et al., 1999; Uehara et al.,

2001) are newly factors identified to be associated with RM and could, in some cases, be

the cause for unexplained RM. However, additional genetic factors contributing to the

aetiology of RM are likely to be detected.

There are different mechanisms by which genetic factors may contribute to

miscarriage, and different modes of inheritance by which these mutations and

polymorphisms are transmitted from parents to foetus. Dominant de novo mutations

arising in the germ cells of the parents may result in affected foetuses and cause

miscarriage. Alternatively, de novo mutations during an early developmental stage in the

foetus may result in mosaicism, which also may cause a severe phenotype resulting in

miscarriage (McFadden and Friedman, 1997; Los et al., 2004). De novo mutations are

difficult to study because these mutations cannot be detected by studying DNA samples

from parents and foetal/placental samples are rarely available.

Autosomal recessive mutations cause a number of genetic diseases, and are common

especially in isolated populations (Peltonen et al., 1999; Norio, 2003a; Teebi and Teebi,

2005; Weinstein, 2007). The role of recessive mutations in specific genes in RM can be

studied without foetal samples by screening the parents for heterozygous mutations, which

when inherited by the foetus in a homozygous state are expected to be lethal. In isolated

populations disease causing mutations are often enriched and may also show uneven

geographical distribution. Evidence of such geographical clustering can be studied using

genealogic data (Kere, 2001; Norio, 2003a).

Not only autosomal mutations, but also abnormalities in the sex chromosomes can

underlie RM. X-linked mutations are predicted to be capable of inducing foetal loss. Some

X-linked recessive mutations are expected to be lethal to all male foetuses inheriting the

mutation and thereby result in miscarriage (Lanasa et al., 1999). Recently, a study

Page 27: Genetic Studies on Recurrent Miscarriage

27

suggested an association between Y-chromosome microdeletions and RM (Dewan et al.,

2006). The role of these genetic factors in the aetiology of RM can be further studied

using parental DNA samples.

In addition to the nuclear genome, the mitochondria contain their own DNA, which is

transmitted only from the mother to the foetus. The human mitochondrial genome is

16,569 bp in size and encodes 37 genes (DiMauro, 2001). Mutations in these genes

leading to dysfunctional proteins may have a role in RM as the mitochondria are involved

in processes necessary for early development (Van Blerkom, 2004).

It is also possible that mutations in imprinted genes can cause miscarriage. When the

imprinting mechanism is malfunctioning loss of function or inappropriate expression can

occur in intact genes and may result in miscarriage (Reik and Dean, 2001; Swales and

Spears, 2005). The role of imprinted genes in RM is, however, difficult to study as

imprinting still is a poorly understood phenomenon.

Recently, several genes such as CTL-A, IL6, ITGB3, and VEGF have been shown to be

associated with RM (Wang et al., 2005; von Linsingen et al., 2005; Ruzzi et al., 2005;

Coulam and Jeyendran, 2008) and it is likely that other genes, in which mutations and

polymorphisms cause or increase the risk for miscarriage, will be identified in the near

future. Therefore, further studies and screening of candidate genes are required to establish

the role of genetic factors in RM.

3.1 Identification of disease genes

3.1.1 The Human Genome Project

The human genome consists of approximately 3 billion base pairs (bp) and is estimated to

contain 20 000-25 000 genes (Balakrishnan et al., 2005; Antonarakis and Beckmann,

2006). The Human Genome Project (HGP) began in 1990 and its main goals were to

determine the nucleotide sequence of the human genome and to identify all human genes.

A draft sequence of the human genome was published in 2001 (Lander et al., 2001; Venter

et al., 2001). An updated sequence, published in 2003, covered ~99% of the euchromatic

genome (International Human Genome Sequencing Consortium 2004). The first complete

Page 28: Genetic Studies on Recurrent Miscarriage

28

genome sequence of a single human individual was published in 2007 (Levy et al. 2007).

In addition to the human genome, the genoms of hundreds of other organisms have been

sequenced (www.ebi.ac.uk/genomes/index.html). This genomic information has

dramatically changed the process of identifying disease genes. It provides new, invaluable

tools for understanding the basic human genetic make-up and how variations in the

genomic sequence result in disease (Peltonen and McKusick, 2001;

http://genomics.energy.gov).

3.1.2 Traditional identification of disease genes

Before the complete human genome sequence was available, disease genes were

traditionally identified using functional and positional cloning methods. In functional

cloning the gene is isolated based on information regarding the protein or its function. The

method can be used if the amino acid sequence of a protein is at least partially known

(Collins, 1992). However, most often, the protein encoded by the disease gene is not

known and the positional cloning method is used to identify a gene based on the

knowledge about its physical location in the genome. Using polymorphic markers and

DNA samples from families with affected members the disease gene locus is localized to a

genomic region by linkage analysis. Disease gene identification based on positional

cloning has been widely used in the research of monogenic diseases (Collins, 1992;

Collins, 1995).

3.1.3 The candidate gene approach

The candidate gene approach is an important step towards disease gene identification and

relies on information about known genes. New effective methods for mutation screening

together with the information offered by the HGP have made the candidate gene approach

a commonly used method to search for disease genes as information on physical locations

and sequences of many genes is available. The candidate gene approach can be used when

the biochemical background of the defect of interest is known, when a chromosomal

region has been linked to a disease or if an animal model of a disease has been established

(Collins, 1992; Ballabio, 1993; Zhu and Zhao, 2007). The mouse is a widely used model

organism for studying mammalian gene function and genes causing a phenotype in the

mouse can be used as candidate genes in human disease studies (Strachan and Read,

Page 29: Genetic Studies on Recurrent Miscarriage

29

2004). Interaction partners of identified proteins defective in a disease can also be

considered as candidate genes for the disease of interest. In addition, candidate genes can

be selected based on results from genome wide expression profiling data showing

differences in gene expression between cells or tissues from affected and control

individuals (Antonarakis and Beckmann, 2006).

3.1.4 Mutation screening

When a candidate gene is selected, the next step in disease gene identification is searching

for mutations in the gene under study. During the past years several different methods

have been used for mutation screening. To date, one of the most commonly used method

is direct sequencing. In addition, denaturing high-performance liquid chromatography

(dHPLC) (Xiao and Oefner 2001) and multiplex ligation-dependent probe amplification

(MLPA) (den Dunnen and White, 2006) are frequently used to screen DNA samples for

sequence variations due to their sensitivity and specificity. A set of patients and control

DNA samples are screened for mutations segregating with the disease phenotype. It is

important to screen a set of healthy control individuals to be able to assess the significance

of the nucleotide changes and to confirm that the variation truly is associated with the

disease (Strachan and Read, 2004).

When screening candidate genes different types of sequence variations can be

identified. Nonsense mutations are point mutations that introduce premature stop codons

into the coding regions. Nonsense mutations usually lead to the degradation of the

transcript or may produce truncated nonfunctional protein products. Missense mutations

are point mutations that result in the substitution of an amino acid and are sometimes

difficult to differentiate from rare neutral polymorphisms. A missense mutation is more

likely to be pathogenic if it changes an amino acid that is conserved. If the change alters

the chemical nature of the side chain of the amino acid this usually suggests a pathogenic

role for the variation. In addition, sequence variations in parts of the gene that code for

functionally important domains are more likely to be pathogenic. Deletions, insertions,

and duplications can disrupt the reading frame, resulting in a completely different protein

translation. In addition to frameshift mutations, deletions, insertions, and duplications

can introduce premature stop codons or alter the protein by amino acid(s) deletion or

insertion (Strachan and Read, 2004).

Page 30: Genetic Studies on Recurrent Miscarriage

30

Splicing mutations affect the consensus sequences at splice donor, splice acceptor, or

splice branch sites and may abolish the splicing partially or completely. In addition,

cryptic splice sites can be activated or splicing enhancers and silencers altered. These

types of splicing mutations may lead to complete or partial exon skipping or intron

retention (Cartegni et al. 2003; Baralle and Baralle, 2005). In addition, neutral and silent

mutations/polymorphisms are likely to be observed in candidate gene screening. A

neutral mutation is a variation that results in the substitution of a different, but

chemically similar amino acid. Silent mutations are variations that do not result in a

change to the amino acid sequence. They may occur either within an exon or within the

non-coding regions of the gene (Strachan and Read, 2004). Polymorphisms in the non-

coding regions of the genome, including the promoter region, the 5´- and 3´-untranslated

regions, and the intragenic regions are also commonly detected. Some of these variations

may affect gene splicing, and transcription and translation, and thereby the expression of

the gene (Wang et al., 2006).

3.1.5 Confirming the nature of a new genetic variation

Identifying a disease gene is a possible starting point to study the function of the protein

encoded by the gene. A candidate disease gene has to be carefully studied to determine if

there is enough evidence that the identified changes are not just neutral polymorphisms

but do cause the disease under study. The candidate gene can be demonstrated to be the

disease gene in different ways. Mutation screening is usually the first step in the process.

Identification of mutations in several unrelated individuals strongly suggests that the

candidate gene is causing the disease, but formal proof requires additional evidence. To

prove that the candidate gene has a pathogenic role in the disease functional studies, such

as restoration of the normal phenotype in vitro, or production of an animal model for the

disease, are required (Strachan and Read, 2004).

3.2 Mouse as a model organism

Genetic approaches in model organisms provide a powerful means by which to examine

the biological basis of human diseases. Over the past century, the mouse has become the

most used animal model system for genetic research because of its close genetic and

Page 31: Genetic Studies on Recurrent Miscarriage

31

physiological similarities to humans. In addition, its small size and short generation time

have allowed large scale mutagenesis experiments and extensive genetic crosses. Many

mouse mutations have been established as reliable models of human disease while others

have provided insight into complex developmental and physiological pathways in

mammals. Targeted mutagenesis approaches in the mouse have led to dramatic increases

in our understanding of human disease processes (Brown, 1998; Brown and Nolan, 1998;

Nolan et al., 2002). There are a number of mouse models described in the literature and

databases on different mutant strains are available on the internet

(http://ceolas.org/VL/mo/; Peters et al., 2007).

Sequencing of the mouse genome was initiated in 1999 and the first draft sequence

was published in 2002 by Waterston and coworkers. Because of the comparatively high

level of sequence conservation between human and mouse coding sequences, almost all

human genes have an easily identifiable mouse homolog. Therefore, disease genes

identified in mouse models are usually good candidate genes for human diseases. The

mouse and human genomes both appear to contain 20 000-25 000 protein coding genes

and more than 90% of the mouse genome can be aligned with a region of the human

genome due to conserved synteny (Waterston et al., 2002). In addition to providing

important data on mouse genes and the structure of the mouse genome, the sequence data

have also been valuable in comparative genomics. One application of human-mouse

comparative sequence analysis involves the identification of orthologous genes.

Functionally important genomic regions are commonly homologous in human and mouse

genomes and genetic information obtained from the mouse can thereby be used for human

disease studies (Collins et al., 1998; Nobrega and Pennacchio, 2004).

Apart from the very early stages of preimplantation development, human embryos are

inaccessible for research and studies are limited to diseased aborted foetuses. Due to the

difficulties of studying implantation and foetal development in humans, much of our

knowledge of the genes involved in these processes has been derived from animal models.

There are several animal models used to gain information about human development

including; nematode worm, fruit fly, rat, guinea pig, rabbit, pig, sheep, cow, and non-

human primates (Lee and DeMayo, 2004; Dvash et al., 2006; Carter, 2007). The most

commonly used model organism is the mouse, but there are, however, significant

differences in the mode of implantation, placental structure, yolk sac, number and function

of placental hormones, and particularly in the length of pregnancy between mouse and

Page 32: Genetic Studies on Recurrent Miscarriage

32

human (Rossant and Cross, 2001; Carter, 2007). Despite these differences, gene ablation

in the mouse has proven to be a powerful tool in elucidating gene function during

embryonic development (Rossant and Cross, 2001; Dvash et al., 2006).

Knock-out mouse models have been used to identify a number of genes required for

normal implantation and placentation (Rossant and Cross, 2001;Lee and DeMayo, 2004;

Carter, 2007). In addition, imprinted genes required for normal embryo development have

been studied in mouse models (Mutter, 1997; Lucifero et al., 2004). A number of mouse

models, in which deletions of a specific gene are embryo-lethal at a specific

developmental stage, have also been described (Healy et al., 1995; Sah et al., 1995; Wang

et al., 1996; Gu et al., 2002). These mouse models can be used to identify candidate genes

for human RM. A gene identified in mouse to be embryo lethal can be considered as a

clinically important candidate gene for human RM if the mouse and human genes have a

highly homologous sequence and similar normal functions in both species.

3.3 Candidate genes for RM

The following sections describe four candidate genes for human RM. They were selected

based on results from studies on animal models. In addition, the genes encoded by the

mitochondrial genome were considered as candidate genes because the mitochondria are

involved in processes important for development but their role in early development is still

unknown.

3.3.1 Amnionless

The amnionless mouse model

The amnionless mouse has a recessive prenatally lethal insertional mutation in the Amn

gene (Wang et al., 1996). The gene was named Amnionless (Amn) as the first visibly

striking feature of the mutant mouse was the lack of an amnion surrounding the foetus.

Examination of the mutant embryos at different times of gestation indicated that the

mutation causes a gastrulation stage (embryonic day (E) 6.5-E8.0) defect in the

homozygous mutant mouse (Wang et al., 1996; Tomihara-Newberger et al., 1998).

Although gastrulation begins normally in the Amn mutants they remain small and

undifferentiated and generate no amnion. The growth and differentiation of the embryonic

Page 33: Genetic Studies on Recurrent Miscarriage

33

ectoderm cells are severely impaired in the mutant embryos. The first signs of abnormal

development in embryos homozygous for the transgene insertion can be seen between

E6.5 and E7.0 and the mutant embryos die and are resorbed between E9.5 and E10.5

(Table 2) (Wang et al., 1996).

Functions of the Amn gene

The Amn gene encodes a type I transmembrane protein (Kalantry et al., 2001). During the

early

post-implantation stages mouse Amn is expressed exclusively in the visceral

endoderm (Tomihara-Newberger et al., 1998; Kalantry et al., 2001). In adults, both mouse

and human Amn is expressed mainly in the kidney and the intestinal epithelium (Tanner et

al., 2003; Strope et al., 2004).

The exact role of the Amn protein in mouse development has not yet been determined,

but the gene function is required in the visceral endoderm for normal primitive streak

formation (Tomihara-Newberger et al., 1998). The visceral endoderm is an epithelial cell

layer that does not contribute directly to the foetus, but is needed to absorb

and digest

nutrients from the maternal environment, to provide signals required for the organization

of the body axes, and to secrete factors that promote correct positioning of the

primitive

streak (Beddington and Robertson, 1999; Martinez-Barbera and Beddington, 2001; Perea-

Gomez et al., 2002). At the onset of gastrulation the middle region of the primitive streak

is absent in the Amn mutant mouse (Tomihara-Newberger et al., 1998).

In wild-type mouse embryos Amn co-localizes with two other gene products, cubilin

(Cubn) and megalin, at the surface of the visceral endoderm. In the Amn mutant mouse

Cubn fails to localize to the surface of the visceral endoderm. This indicates that Amn is

an essential component of the Cubn receptor complex, and it has been suggested that the

Amn/Cubn-complex is required for endocytosis/transcytosis of one or more ligands

in the

visceral endoderm. (Strope et al., 2004). In humans, AMN and CUBN also form a

complex. This complex is required for uptake of vitamin B12 (Fyfe et al., 2004; Luder et

al., 2008). Mutations in either of these genes cause selective intestinal malabsorption of

vitamin B12, a rare autosomal recessive disorder known as Imerslund-Gräsbeck syndrome

(IGS) or megaloblastic anemia 1 (Imerslund, 1960; Gräsbeck et al,. 1960; Aminoff et al.,

1999; Tanner et al., 2004).

Page 34: Genetic Studies on Recurrent Miscarriage

34

The fact that the Amn protein has a similar structure in both mice and humans suggests

similar functions in both species, making Amn a candidate gene for miscarriages in

humans. Amn may be required in the extra-embryonic endoderm during early embryonic

development, and in the intestine for the uptake of vitamin B12 in adults in both species.

Table 2. Comparison between mouse and human stages of early development and the timepoints at which

TM, AMN, and EPCR mutant mice embryos die. The carnegie stages (CS) provide a universal system for

staging and comparing the embryonic development of vertebrates. The stages are based on the external and

internal morphological development of the embryo, and are not directly dependent on either age or size. The

stages covers approximately the first 60 days of human development and nearly the entire intrauterine

development of mouse (UNSW Embryology, http://embryology.med.unsw.edu.au/embryo.htm) The table

shows the age of human and mouse embryos in days at specific stages. Implantation takes place during CS

5, gastrulation starts during CS 7, neurogenesis starts during CS 8, and heart beats can be detected at CS 10.

Developmental stages at

Carnegie stage Human Age Mouse Age lethality of mutations

CS 1 E 1 E 0.5

CS 2 E 2-3 E 1-2

CS 3 E 4 E 3

CS 4 E 5-6 E 4

CS 5 E 7-12 E 5-6 Implantation

CS 6 E 13-15 E 6.5

CS 7 E 15-17 E 7 Gastrulation

CS 8 E 17-19 E 7.5 Neurogenesis

CS 9 E 19-21 E 8

CS 10 E 22-23 E 8.5 Heart beats

CS 11 E 23-26 E 9 TM mutants die

CS 12 E 26-30 E 9.5

CS 13 E 28-32 E 10 AMN mutants die

CS 14 E 31-35 E 10.5 EPCR mutants die

CS 15 E 35-38 E 11

CS 16 E 37-42 E 11.5

CS 17 E 42-44 E 12

CS 18 E 44-48 E 13

CS 19 E 48-51 E 13.5

CS 20 E 51-53 E 14.5

CS 21 E 53-54 E 15

CS 22 E 54-56 E 15.5

CS 23 E 56-60 E 16

Page 35: Genetic Studies on Recurrent Miscarriage

35

3.3.2 Thrombomodulin and Endothelial Protein C Receptor

The Thrombomodulin mouse model

In mice, absence of the blood-clotting regulator thrombomodulin (TM) causes embryonic

lethality before development of a functional cardiovascular system (Healy et al., 1995).

TM has been shown to have a dual role in development as TM is needed in two distinct

tissues. Expression of TM in non-endothelial cells of the placenta is required for proper

function of the early placenta, while the absence of TM from blood vessel endothelium

causes excessive activation of the embryonic blood coagulation system (Healy et al., 1995;

Isermann et al., 2001). The first indications of abnormal development are observed at E8.5

in homozygous TM-deficient (TM-/-) embryos and manifest as overall growth retardation.

This initial phenotype is invariably followed by rapid and complete resorption of

homozygous TM-null embryos within the next 10-12 hours (Table 2) (Healy et al., 1995;

Isermann et al., 2001).

Functions of the TM gene

TM is an endothelial cell surface glycoprotein expressed on the endothelium of arteries,

veins, capillaries, lymphatics, and on trophoblast cells of the placenta (Maruyama et al.,

1985; Weiler-Guettler et al., 1996; Fazel et al., 1998). TM plays a key role in the protein C

anticoagulant pathway, which is a major regulatory system providing an inhibitory

feedback mechanism that suppresses blood coagulation by preventing fibrin formation.

TM inhibits blood coagulation by forming a 1:1 complex with thrombin. Binding of

thrombin to this high-affinity receptor alters its specificity toward several substrates. The

complex converts protein C into the natural anticoagulant activated protein C (APC)

approximately 1000 times faster than thrombin alone. APC proteolytically inhibits the

coagulation cofactors Va and VIIIa, resulting in down-regulation of the activity of the

coagulation system (Figure 3) (Maruyama et al., 1985; Van de Wouwer et al., 2004;

Dahlbäck and Villoutreix, 2005).

A number of polymorphisms in the TM gene have been characterized and their

association with thrombosis or arteriosclerosis are subjects of ongoing research. These

associations have been evaluated in several studies and the results suggest that the

relevance of TM mutations in these complications is small or limited to a small subgroup

of individuals (Ireland et al., 1997; Doggen et al., 1998; Le Flem et al., 1999; Nakazawa et

al., 2002). The most convincing evidence that reduced TM function causes thrombosis is

Page 36: Genetic Studies on Recurrent Miscarriage

36

derived from animal studies. In the past three genetically engineered mouse lines with various degrees of TM dysfunction have been generated, showing different degrees of thrombosis (Healy et al., 1995; Healy et al., 1998; Isermann et al., 2001). It appears that mutations causing a substantial reduction of TM function are very rare in humans (Faioni et al., 2002), possibly because such loss-of-function mutations would lead to foetal loss at an early developmental stage.

Figure 3. The protein C anticoagulant pathway. 1) Activation of prothrombin (PT); the Xa-Va complex converts prothrombin to thrombin (T), which forms a complex with TM. 2) Protein C activation; protein C (PC) binds to EPCR, which presents PC to the T-TM complex. 3) Cofactor inactivation; activated protein C (APC) binds protein S (PS). This complex inactivates factors Va and VIII thereby preventing blood coagulation. The EPCR mouse model Deletion of the endothelial protein C receptor (EPCR) gene in mice leads to embryonic lethality. Genotyping of progeny obtained from EPCR+/+/ interbreeding indicates that EPCREPCR / embryos die on or before E10.5 (Table 2). However, EPCRHowever, EPCR // embryos removed from extra-embryonic membranes and tissues at E7.5 and cultured in vitro develop beyond E10.5, suggesting a role for EPCR in the normal function of the placenta and/or at the maternal-embryonic interface (Gu et al., 2002). EPCR is normally detected on giant trophoblast cells, which are in direct contact with the maternal circulation and its clotting factors. In the giant trophoblast cells derived from EPCRcells derived from EPCR // embryos thrombosis is observed (Li et al., 2005). These observations provide evidence that extraembryonic EPCR expression is essential for embryonic viability and plays a critical role in the control of blood coagulation at the foeto-maternal boundary (Gu et al., 2002).

TM

Page 37: Genetic Studies on Recurrent Miscarriage

37

Functions of the EPCR gene

EPCR is a type 1 transmembrane protein and is strongly expressed in the endothelial cells

of arteries and veins in heart and lung, and less intensely expressed in capillaries in the

lung and skin (Laszik et al., 1997; Simmonds and Lane, 1999; Esmon, 2000). It is also

expressed in the placenta and the developing cardiovascular system of the foetus (Crawley

et al., 2002). EPCR is seen as a complementary cofactor in protein C activation (Figure 3).

It binds both protein C and APC with similar affinity and functions in the protein C

pathway by binding protein C and presenting it to the TM-thrombin complex on the

endothelium, thereby further enhancing the rate of protein C activation and suppressing

blood coagulation (Stearns-Kurosawa et al., 1996; Van de Wouwer et al., 2004).

In human, sequence variations in the EPCR gene are associated with the risk of

thrombosis. A heterozygous mutation in EPCR has been identified in humans who

develop thrombosis and myocardial infarction. This mutation, c.323-9_336dup, is a 23-bp

insertion that leads to the formation of a truncated receptor that lacks the extracellular and

transmembrane domains and is unable to sustain protein C activation (Biguzzi et al.,

2001). However, a common polymorphism in EPCR, c.717+16G>C, seems to reduce the

risk of thrombosis. The CC genotype reduces the risk of venous thromboembolism 2.6-

fold, probably due to the higher APC levels observed in individuals carrying this genotype

(Medina et al., 2004; Medina et al., 2005).

TM and EPCR as candidate genes

Pregnancy complications in patients with prothrombotic risk factors have mainly been

linked to the formation of blood clots in placental blood vessels and sinuses, eventually

leading to intrauterine growth retardation and, in the most severe case, foetal loss

(Kupferminc, 2003; Kujovich, 2004; Kutteh and Triplett, 2006). Therefore, mutations in

the genes involved in the blood coagulation systems may increase the risk of thrombosis

and thereby miscarriage. Although the importance of the TM-protein C-EPCR system for

placental development and maintenance of pregnancy is firmly established in mice (Healy

et al., 1995; Gu et al., 2002), the relevance of these mechanisms for pregnancy associated

complications in humans remains to be established. The data from mouse models, the

known sequence homology of mouse and human TM and EPCR, and the similar type of

placentation in both species (Dittman and Majerus, 1989; Cross et al., 1994), however,

suggests TM and EPCR as candidate genes for RM in humans.

Page 38: Genetic Studies on Recurrent Miscarriage

38

3.3.3 p53

Functions of the p53 gene

The p53 gene encodes a transcription factor consisting of an N-terminal transactivation

domain, a central sequence specific DNA binding domain, and a C-terminal

oligomerization domain (Levine, 1997). p53 is expressed at low levels in most cells and

the protein levels increase in response to various stress signals. p53 is expressed at high

levels in the mouse embryo up to midgestation, after which the expression levels rapidly

decrease (Rogel et al., 1985; Marzusch et al., 1995). Hundreds, if not thousands, of studies

have been dedicated to the functions of p53 because it plays a pivotal role in tumour

suppression. Indeed, more than 50% of human cancers contain somatic mutations in this

gene (Levine, 1997). p53 functions as a tumour suppressor by regulating the cell cycle,

protecting the genome from DNA damage and in inducing apoptosis. Moreover, p53

appears to interact with a large number of proteins involved in growth control, DNA repair

and transcriptional regulation (Choi and Donehower, 1999; Cheah and Looi, 2001; Balint

and Vousden, 2001). Experiments in mouse models have revealed that most of the human

tumour suppressor genes are essential for embryonic development. Furthermore,

inactivation of both germ line alleles of a tumour suppressor usually results in embryonic

lethality either in early or midgestation (Rogel et al., 1985; Schmid et al., 1991; Jacks,

1996), making p53 a candidate gene for RM.

There is one previous study supporting the hypothesis that p53 has a potential role

during early pregnancy also in humans. The results of Pietrowski et al. (2005) indicate that

RM is associated with a polymorphism 12139G>C at codon 72, resulting in either a

proline or arginine residue. In addition, the frequency of the Pro72 allele has been shown

to be significantly higher among women experiencing recurrent implantation failure than

in control women (Kay et al., 2006). The codon 72 variation of p53 have been shown to

differ biochemically and biologically (Thomas et al., 1999), but the role of p53 and the

codon 72 polymorphism in human development needs to be further studied.

p53 mouse models

Although p53 null mice can be viable and fertile, the absence of p53 activity can lead to

fatal developmental abnormalities and a significant proportion of p53-/- mice die during

embryogenesis. A substantial fraction of p53 null mice have developmental defects,

including neural tube defects, and the affected embryos frequently undergo resorption

Page 39: Genetic Studies on Recurrent Miscarriage

39

clearly implicating p53 as an important factor in the early development (Sah et al., 1995;

Armstrong et al., 1995).

p53 can inhibit DNA replication by triggering growth arrest in G1 or initiate apoptosis,

and therefore its activity must be tightly regulated in order to allow normal growth. In the

absence of stress inducing signals, p53 is kept at low levels via its interaction with mouse

double minute 2 (MDM2) that binds to the transcriptional activation domain of p53. For

cell division in the early embryo to proceed, p53 needs to be down-regulated by Mdm2

(Piette et al., 1997; Bond et al., 2005). A deletion of Mdm2 causes embryonic lethality in

mice around the time of implantation (Montes de Oca Luna et al., 1995; Jones et al.,

1995). In addition, p53QS

mutant mice, which cannot bind Mdm2 due to mutations in

residues crucial for transactivation, show increased embryonic lethality in a homozygous

state (Johnson et al., 2005).

Mouse models have also suggested that p53 may play a key role in development

through protecting embryos from teratogens and diverse environmental stresses (Nicol et

al., 1995; Norimura et al., 1996). Teratogenesis experiments in mice provide strong

support for the idea that p53 efficiently aborts embryonic cells with teratogen induced

DNA damage. If too many cells of the embryo die through apoptosis, the entire embryo

will die. Conversely, lack of p53 in the embryo presumably results in continued cell cycle

progression despite DNA damage, much less apoptosis and a higher proportion of

embryos with developmental anomalies (Nicol et al., 1995). Taken together, the results

obtained from mouse models suggest that the relative levels of p53 in the early embryo are

highly important, and the protein expression has to be very carefully controlled during

developmental processes. Changes in the expression levels of p53 elevates the risk of

miscarriage and developmental defects, and increases the susceptibility to chemical

teratogenesis, possibly also in humans.

3.3.4 The mitochondrial genome

Mitochondria are cellular organelles that contain their own genome. It is estimated that in

a somatic cell one human mitochondrion typically contains 5-10 DNA copies and,

depending on the cell type, a somatic cell can contain up to 1,000 mitochondria. The

human mitochondrial genome is circular, double-stranded and 16,569 bp in size. It

Page 40: Genetic Studies on Recurrent Miscarriage

40

contains 37 genes; 13 protein coding genes, which are all components of the respiratory

chain; two rRNAs; and 22 tRNAs needed for mitochondrial protein synthesis (Howell et

al., 2000; DiMauro, 2001; Thorburn and Dahl, 2001). The mitochondria are involved in

cellular ATP production and apoptosis (DiMauro, 2007). These processes are of great

importance in the early development of embryos and may be disturbed by mutations in

mitochondrial DNA (mtDNA).

Genetics of mtDNA

Mitochondrial genetics differs from Mendelian genetics in three major aspects:

1) Maternal inheritance. At fertilization, all mitochondria in the zygote derive from the

oocyte. mtDNA mutations are therefore inherited in a matrilineal fashion (DiMauro, 2001;

McKenzie et al., 2004).

2) Heteroplasmy/threshold effect. In contrast to autosomal nuclear genes, each

consisting of one maternal and one paternal allele, there are hundreds or thousands of

mtDNA molecules in each cell. Deleterious mutations in mtDNA usually affect some but

not all of the mtDNA genomes. Therefore cells, tissues, and organs can contain two

different mtDNA populations (wild-type and mutant), a state known as heteroplasmy. The

situation when all mtDNAs are identical is called homoplasmy. Non-deleterious

mutations/neutral polymorphisms of mtDNA are usually homoplasmic, whereas

pathogenic mutations are usually heteroplasmic. For an mtDNA mutation to have a

deleterious effect on tissue function it must reach a minimum threshold - a certain

proportion of mutant molecules are required before a phenotype is observed (DiMauro,

2001; Thorburn and Dahl, 2001; McKenzie et al., 2004).

3) Segregation. As a heteroplasmic cell replicates different proportions of mutant

mtDNA molecules are transmitted to the daughter cells, and when the pathogenic

threshold for a given tissue is surpassed the phenotype can also change. In this way the

mtDNA genotype can drift toward homoplasmy, either wild-type or mutant (DiMauro,

2001; McKenzie et al., 2004).

Today over 200 mtDNA point mutations have been reported in the Mitomap database

(http://www.mitomap.org) to be associated with different human diseases. These

mutations can be divided into two groups: mutations affecting mitochondrial protein

synthesis, including mutations in tRNA and rRNA genes and mutations in protein

encoding genes. The mutations are involved in a broad spectrum of human diseases,

Page 41: Genetic Studies on Recurrent Miscarriage

41

including multisystem disorders as well as more tissue specific diseases (DiMauro, 2001;

Taylor and Turnbull, 2005; McKenzie et al., 2004; Wong, 2007).

The mitochondrial bottleneck

The segregation of mtDNA alleles is unexpectedly rapid, and in some maternal lineages

complete allele switching has been observed in a single generation without a

heteroplasmic intermediate (Koehler et al, 1991). Rapid changes in mitochondrial allele

frequencies were first observed in cattle (Olivo et al., 1983; Ashley et al., 1989) but

similar changes have subsequently been described in many mammalian species, including

humans. Analysis of human pedigrees segregating pathogenic mtDNA point mutations

show large differences in the proportion of mutant mtDNAs among siblings and between

generations (Larsson et al., 1992; Ghosh et al., 1996; Jacobi et al., 2001).

The rapid changes in mtDNA allele frequency have been explained by an mtDNA

bottleneck in the germ line, responsible for the random segregation of mtDNA sequence

variations between generations. A mature oocyte contains about 100,000-200,000 copies

of mtDNA (Figure 4). These mtDNA copies are randomly divided, without mtDNA

replication, into cells of the developing embryo. Replication of mtDNA is initiated when

the primordial germ cells (PGC) migrate and differentiate to generate oocytes. It is

estimated that human PGCs contain ~10-100 mitochondria, increasing to ~200-1000 in the

oogonia and to several thousands in the oocyte in the primordial follicle. As the oocyte

matures the number of mitochondria again increases to >100,000 (Jenuth et al., 1996;

Shoubridge, 2000; Khrapko, 2008; Stewart et al., 2008). Segregation is rapid because the

precursor cells contain relatively small numbers of mtDNA templates and because the

replication of mtDNA is under relaxed control. This relaxed form of replication allows

some templates to replicate more than once during the cell cycle, while others are not

replicating at all. Because of the mitochondrial bottleneck, the mtDNA genotype in the

embryo is largely determined by the genotype of the few mitochondria in the PGC in the

mother (Jenuth et al., 1996; Shoubridge, 2000).

mtDNA in development

The mitochondria produce the energy needed for cellular processes by producing ATP

through oxidative phosphorylation. Through dysfunctions of energy metabolism or

regulation of apoptosis, the mitochondria are suggested to contribute to human oocyte

wastage and early embryo demise (Van Blerkom, 2004). The role of mtDNA mutations in

Page 42: Genetic Studies on Recurrent Miscarriage

42

human early development is supported by a report describing a family with an unusual

homoplasmic tRNA mutation that resulted in six neonatal deaths and one surviving child

with Leigh syndrome even though the mother is clinically unaffected (McFarland et al.,

2002). As the proportion of mutant mtDNA copies can shift when inherited the phenotype

often shows remarkable variation within a family (Ghosh et al., 1996; Blok et al., 1997;

Jacobi et al., 2001). Accordingly, such a shift can occur from a mother with mild or no

phenotype to an embryo with significant enrichment of mutated mtDNA causing foetal

loss and possibly RM. Homoplasmy for a severe pathogenic mtDNA mutation is rarely

observed, presumably because such a condition would be lethal. For example mutations

causing mitochondrial encephalopathy with lactic acidosis and stroke-like episodes

(MELAS) and myoclonic epilepsy and ragged-red fibres (MERRF) are invariably

heteroplasmic and homoplasmic mutations are presumed to be fatal (Jenuth et al., 1996;

Howell et al., 2000). An oocyte with such a mutation in a homoplasmic state would most

likely result in foetal death at an early developmental stage. It has been suggested that this

phenomenon might even be beneficial, if spreading of the deleterious mtDNA mutations

that slip through the bottleneck would thus be eliminated (Bergstrom and Pritchard, 1998;

Howell et al., 2000).

Figure 4. The mitochondrial bottleneck. During oocyte maturation a limited number of mtDNA copies are

transferred to each primary oocyte. As this small population of mtDNA copies is rapidly replicated it may

lead to a random shift of the mutational load in the next generation. Due to the mitochondrial bottleneck the

offspring of a heteroplasmic female may receive a varying number of mutant mtDNA copies at fertilization

resulting in different phenotypes; A) low level of mutation (unaffected offspring), B) intermediate level of

mutation, C) high level of mutation (affected offspring). Modified from Taylor and Turnbull, 2005.

Page 43: Genetic Studies on Recurrent Miscarriage

43

Aims

The present study is based on hypotheses that some as yet unknown genetic factors may

result in RM. Consequently, the main aim of this study was to gain new information about

the underlying genetic causes of RM in the Finnish population. The aim was to identify

candidate genes for RM and subsequently screen these genes for mutations in a series of

samples with RM. In addition, we aimed to study if genetic factors previously reported to

underlie RM could be the cause of RM in Finnish patients.

The following specific aims were set for the study:

1. To determine whether ancestral places of origin of the couples with RM show

geographic clustering.

2. To search for mutations in four candidate genes for RM ─ Amnionless,

Thrombomodulin, Endothelial Protein C Receptor and p53 ─ identified to be cruical for

development in mouse models.

3. To study the expression of p53 in placental samples.

4. To search for variations associated with RM in the mitochondrial genome.

5. To study if specific sex chromosome characteristics, namely skewed X chromosome

inactivation or Y chromosome microdeletions, are associated with RM in Finnish patients.

Page 44: Genetic Studies on Recurrent Miscarriage

44

Materials and methods

1 Study subjects

1.1 Patients (I, II, III, IV, V)

The patients recruited for this study have been treated for unexplained RM at the

Department of Gynaecology and Obstetrics of the Helsinki University Hospital during the

years 2001-2004. The inclusion criteria for the 48 women included in the study were:

1) age 18-40 years (mean age 31 years)

2) previous history of RM, defined as three or more consecutive miscarriages

3) normal karyotype

4) no uterine abnormalities (examined by ultrasonography or hysterosonogram)

Blood samples were obtained from male partners of 40 of the women included in the

study. The karyotypes of these males were checked and found to be normal. Consequently,

a total of 40 couples and 8 women with unexplained RM were included in the study. The

women in the study had experienced three to seven miscarriages. In 45 women all

miscarriages had taken place during the first trimester (weeks 1-12 of pregnancy). Two

women had experienced a second trimester (weeks 13-26 of pregnancy) miscarriage, and

one women a third trimester (from week 27 to birth) intrauterine foetal death in addition to

first trimester losses.

1.2 Controls (I, II, III, IV, V)

The control group consisted of 191 healthy women who had had at least one normal

pregnancy and no known history of miscarriage. They were recruited from the same

hospital during the same time period as the patients. The controls were of the same age

and racial origin (Finnish Caucasian) as the patients.

Page 45: Genetic Studies on Recurrent Miscarriage

45

1.3 Placental samples (III, IV)

One or more paraffin embedded placental sample from a spontaneously aborted pregnancy

were available from seven of the RM women included in the study. Altogether, placental

samples were available from 19 miscarriages.

1.4 Ethical issues (I, II, III, IV, V)

Informed consent was obtained from all participants prior to enrolment in the study. The

ethics committee of the Department of Obstetrics and Gynaecology, Helsinki University

Central Hospital, has approved this study.

2 Methods

2.1 Genealogic analysis (I)

Data concerning the birthplaces of the patients´ parents and grandparents were collected to

determine the ancestral places of origin for the patients and to investigate if the birthplaces

were clustered to specific regions of Finland. Data concerning the birthplaces were

obtained through a questionnaire. Replies were received from 82 (95%) patients. To

investigate the geographic distribution of ancestral birthplaces, the birthplaces were placed

onto a map of Finland, including the formerly Finnish areas of Russian Karelia and

compared to the population distribution of Finland.

2.2 DNA extraction (I, III)

DNA was extracted according to the manufacturer´s instructions from whole blood

samples and paraffin sections using Puregene DNA isolation kits (Gentra Systems,

Minneapolis, USA).

Page 46: Genetic Studies on Recurrent Miscarriage

46

2.3 Candidate gene analysis

Polymerase chain reaction and dHPLC sample analysis (I, II, III, IV)

The exons and exon-intron boundaries of the candidate genes were amplified using

polymerase chain reactions (PCR). Amplification of appropriately sized PCR products

were confirmed by agarose gel electrophoresis before further analysis. Mutation screening

of the candidate genes was mostly carried out using denaturing high performance liquid

chromatography (dHPLC) with the Transgenomic WAVE Nucleic Acid Fragment

Analysis System (Transgenomic, Omaha, USA) and the associated Navigator software. To

obtain optimal resolution of homoduplex and heteroduplex DNA fragments the

temperature was set for partially denaturing conditions. The melting profile of the

amplicons and the optimal oven temperature was predicted for each PCR fragment using

the Navigator software. The buffer gradients for the elution of the fragments were created

automatically with the Navigator software.

To allow sequence analysis of low-level heteroplasmic mtDNA, fractions of small

sized heteroduplex peaks were collected using an FCW-200 fraction collector integrated

with the WAVE system. To increase the heteroplasmy level in the samples the DNA

fractions were reamplified with the primers used to produce the original amplicon.

Sequencing (I, II, III, IV)

Following dHPLC analysis, samples showing heterozygous peaks were sequenced in order

to determine the nature of the sequence change. PCR products were purified and

sequenced using BigDye version 3.1 sequencing chemistry and an ABI 3730 DNA

Analyzer (Applied Biosystems, Foster City, USA) according to the manufacturer's

instructions.

Genotype determination (I, II, III)

Variations that were detected in a homozygous state by sequencing were genotyped by

restriction enzymes, sequencing, allele specific PCR, or dHPLC.

Solid-phase minisequencing (IV)

Quantification of heteroplasmy levels were performed using solid-phase minisequencing

according to a previously described protocol (Suomalainen and Syvänen, 2000). PCR

Page 47: Genetic Studies on Recurrent Miscarriage

47

products were amplified using specific primers designed for each of the variations studied.

Biotinylated PCR products were then captured in streptavidin-coated microtiter wells and

single nucleotide extension of specific detection primers were performed using

radioactively labelled nucleotides. The results were expressed as numeric cpm-values,

where the incorporated label expressed the relative amount of sequence variation in the

sample.

Quantitative real-time PCR (IV)

The mitochondrial cytochrome b (Cytb) and the nuclear amyloid beta precursor protein

(APP) genes were simultaneously amplified by quantitative TaqMan real-time PCR assay

in an ABI Prism 7000 Detection System Cycler to quantify the amount of mtDNA.

Fluorescence acquisition was done at the end of each annealing step, and the relative

levels of the genes were compared in the exponential phase of PCR, using the ABI Prism

7000 SDS software (Applied Biosystems). The ratio of mtDNA to nuclear DNA was used

as a measure of mtDNA content (2-ΔΔC

T method) (Livak and Schmittgen, 2001).

Predicting the effects of variations (I, II, III, IV)

The ESEfinder (http://exon.cshl.edu/ESE/) is a web based resource which scans nucleotide

sequences to identify putative exonic splicing enhancers (ESEs), which act as binding sites

for splicing factors. The ESEfinder was used to predict if the exonic or intronic sequence

variations disrupt the sequence of known ESE-elements. The ESEfinder searches for

putative ESE motifs and calculates a score for all sequences with a motif match. A

sequence is considered to be potentially needed for splicing when the calculated score is

greater than the threshold value defined by the program.

The variations predicted to change an amino acid were analyzed by the SIFT (sorting

intolerant from tolerant) program (http://blocks.fhcrc.org/sift/SIFT.html) which predicts

whether an amino acid substitution in a protein will be tolerated.

Different web based databases were used to determine if the detected variations have

been previously reported; Ensemble Genome browser (http://www.ensembl.org), UCSC

Genome browser (http://genome.ucsc.edu), NCBI Entrez SNP database

(http://www.ncbi.nlm.nih.gov), p53 knowledgebase (http://p53.bii.a-star.edu.sg),

Mitomap (http://www.mitomap.org), and mtDB-Human Mitochondrial Genome Database

(http://www.genpat.uu.se/mtDB/).

Page 48: Genetic Studies on Recurrent Miscarriage

48

2.4 Immunohistochemical analysis of p53 (III)

5μm-thick sections were cut from paraffin blocks containing placental samples. The

sections were deparaffinized in xylene and rehydrated through graded concentrations of

ethanol to distilled water. Monoclonal antibody DO-7 directed against p53 protein

(dilution 1:100; Dako, Glostrup, Denmark) was used as primary antibody (incubation time

55 min). The procedure was performed in a TechMate 500 automated machine (DAB

detection kit; Dako ChemMate). The sections were lightly counterstained with Mayer's

hematoxylin.

2.5 X chromosome inactivation analysis (V)

The XCI ratio was determined in the females by analysing the methylation status of the X-

linked androgen receptor (AR). Exon 1 of the AR gene contains a polymorphic

microsatellite with varying numbers of trinucleotide repeats (CAG)n which can be used to

identify the different alleles of the locus. DNA was digested with a methylation sensitive

restriction enzyme that digests only nonmethylated (active X) DNA segments. Methylated

sites can not be digested and remain intact for amplification.

Fragment analysis of the microsatellite was performed using an ABI 3730 DNA

Analyzer (Applied Biosystems). Fragment analysis data were analysed and genotyped

using GeneMapper 3.0 software. The total fluorescent peak areas for both alleles in

digested and undigested samples were recorded and used for calculation of the ratio of

allele inactivation. Previously described equations (Lau et al., 1997; Iitsuka et al., 2001)

were used to calculate the degree of inactivation. We used both 85% (mildly skewed) and

90% (extremely skewed) inactivation of a particular X chromosome as cut off points for

skewed X-inactivation.

2.6 Y chromosome microdeletion analysis (V)

Males were screened for the presence of microdeletions using Y chromosome specific

sequence tagged sites (STS) primer sets. Altogether, 37 STS loci spanning the whole Y

chromosome were analysed in multiplex PCR reactions. The presence of the specific Y

chromosome markers were shown by observing the PCR fragments on agarose gels.

Page 49: Genetic Studies on Recurrent Miscarriage

49

2.7 Statistical analysis (I, II, III, IV, V)

X 2 and Fisher´s exact tests ([SISA 1997 programs] http://home.clara.net/sisa) were used

for statistical analyses of the data. The odds ratios (OR) and 95% confidence intervals (CI)

of the OR were calculated to assess the relative risk conferred by a particular allele or

genotype. Fisher´s exact test was also used to compare the frequency of skewed XCI in the

patient and control groups. The T-test was used to calculate differences in mtDNA copy

number levels after real-time PCR analyses. Differences were considered as statistically

significant for p-values <0.05.

Page 50: Genetic Studies on Recurrent Miscarriage

50

Results and discussion

The fact that the observed incidence of RM is much higher than expected by chance alone

suggests that RM is a clinical entity distinct from sporadic miscarriages (Regan, 1991).

This is supported by the finding that RM tends to occur, unlike sporadic miscarriage,

when the foetus has a normal chromosome constitution (Stirrat, 1990; Sullivan et al.,

2004). Recently, the genetic knowledge about the aetiology of diseases has expanded

rapidly. Progress has, however, been much slower in elucidating the genetic causes of

female infertility and RM. Although many studies have been conducted to identify the

underlying mechanisms, the cause of miscarriage remains unknown in approximately 50%

of cases (Plouffe et al., 1992; Clifford et al., 1994). Consequently, this study was

conducted to gain further information about the possible genetic causes of RM in our

Finnish patients.

1 Genealogic studies (I)

Genealogical studies can be used to identify geographic enrichment of genetic disorders.

Depending on the age of the founder mutation the geographical area of such an enrichment

can be very restricted or include large regions. In many diseases of the Finnish disease

heritage a regional clustering of cases can still be observed, reflecting the geographical

origins of ancestral birthplaces (Norio, 2003a; Norio, 2003b). For example, a specific type

of ovarian failure, caused by mutations in the follicle stimulating hormone receptor

(FSHR), called follicle stimulating hormone-resistant ovaries (FSH-RO) is an autosomal

recessive trait which shows geographical enrichment of ancestral birthplaces (Aittomäki et

al., 1995; Aittomäki et al., 1996). We conducted genealogic studies to investigate if such

an enrichment reflecting a common genetic cause could underlie the miscarriages in our

patients. The ancestors of all couples with RM were traced by questionnaires enquiring as

to the full names, dates and places of birth of the patients´ parents and grandparents.

Replies were received from 82 patients. To investigate the geographic distribution of the

birthplaces, the birthplaces of the parents and grandparents were placed onto a map of

Finland (Figure 5) and compared to a reference map showing the distribution of the

Finnish population.

Page 51: Genetic Studies on Recurrent Miscarriage

51

The results of the genealogic studies suggest the possibility of a shared underlying

genetic defect for the miscarriages in a subset of patients. Compared to the reference map

of the general population distribution in Finland the geographical distribution of the

birthplaces of the parents and grandparents of the patients is uneven (Figure 5). The results

suggest that there may be an eastern enrichment in the province of Kuopio of the ancestral

birthplaces. In addition, the Helsinki region is over-represented, but this is explained by

the fact that the patients were all collected from Helsinki. The clustering around Kuopio

suggests that some of the patients may, due to common ancestors, carry a mutation

increasing the susceptibility to RM. This would not be exceptional in Finland due to the

population history, which has increased the local incidence of rare disorders. The special

assortment of rare diseases among Finns is caused by the fact that the population living in

Finland today is descended from a relatively small original population. As the population

has increased, some alleles increased in frequency while others disappeared due to genetic

drift, population bottlenecks, and non-assortative mating (Norio, 2003a; Norio, 2003b).

Figure 5. Birthplaces of the A) parents and B) grandparents of 82 patients with RM. 34 of the parents and

24 of the grandparents were born in Helsinki. The population density of Finland at the time the parents were

born is shown in the small reference map.

A) B)

Page 52: Genetic Studies on Recurrent Miscarriage

52

2 Candidate gene analysis (I, II, III, IV)

The results of the genealogic studies suggest a common underlying genetic cause for RM

in at least a subgroup of our patients. Therefore, we screened a number of candidate genes

for mutations that could possibly explain the RM experienced by the couples. The

candidate genes AMN, TM, EPCR, and p53 were selected based on findings in animal

models. When screening candidate genes in which homozygous mutations are expected to

be lethal in early pregnancy, the affected foetuses cannot, in most cases, be studied

because DNA from the spontaneously aborted foetuses is rarely available. Such mutations

can be studied using placental tissue samples or by screening couples experiencing RM, in

which case partners are expected to be healthy heterozygous carriers of mutations in the

candidate gene of interest. In addition to candidate genes identified from animal models,

the genes of the maternally inherited mitochondrial genome were chosen as candidate

genes based on their known functions.

2.1 Amnionless (I)

Due to the fact that homozygous mutations in Amn are known to cause miscarriage in the

mouse (Wang et al., 1996) and the human AMN and mouse Amn genes are highly

homologous (Kalantry et al., 2001) AMN was considered a candidate gene for

miscarriages in humans. Therefore, we examined whether there is an association between

RM and sequence variations of the AMN gene in humans. To our knowledge this is the

first study concerning the role of AMN in miscarriage.

2.1.1 Variations in the Amnionless gene

A total of 14 sequence variations were found by screening 85 patients (40 couples and 5

women) with a history of unexplained RM and 95 controls for mutations in the 12 exons

and the untranslated regions (UTR) of the Amnionless gene. Eleven of the detected

sequence variations are intronic. Of these variations, 10 are single nucleotide substitutions,

and one variation is a 10 basepair (GCGTGGCGTG) duplication. Three of the sequence

variations are exonic, of which two are single nucleotide substitutions and one is a

duplication (Table 3). None of the exonic variations detected in AMN had been previously

Page 53: Genetic Studies on Recurrent Miscarriage

53

reported (databases used: www.ncbi.nlm.nih.gov; www.ensembl.org). In total, 8 novel

sequence variations were detected in the AMN gene (Table 3).

The sequence variation c.363G>A is a single nucleotide substitution in exon 5,

changing the last nucleotide of codon 121 (GGG->GGA). It is a synonymous variation,

detected only in patient samples and does not predict a change of the amino acid (glycine).

Sequence variation c.829A>G is a non-synonymous SNP in exon 8, changing the first

nucleotide of codon 276 (ACC->GCC). The variation is predicted to change the amino

acid from threonine to alanine. According to the SIFT program this amino acid change

would be tolerated even though the hydroxyl group of threonine makes it much more

hydrophilic and reactive than alanine. Sequence variation c.1339_1344dup is a duplication

in exon 12. It is a 6 basepair duplication (GCCGGG) of the codons 447 and 448, located 5

codons before the stop codon. This DNA sequence change is predicted to make the final

protein product 2 amino acids (Ala+Gly) longer.

Table 3. The location of the variations in the AMN gene, predicted amino acid changes, and the numbers of

patients and controls carrying the rarer allele in a hetero/(homozygous) state. Numbering of the nucleotides

is relative to the adenine in the ATG start codon in the reference sequence (UCSC Genome Browser,

http://genome.ucsc.edu/).

AMN sequence variations

RM women+

DNA variation male partners RM women controls

(predicted amino acid change) Location n=85 n=45 n=95

c.-87C>G Us 5´UTR 13 9 15

c.-74C>T Us 5´UTR 10 8 11

c.-27T>C Us 5´UTR 48 (15) 24(9) 50 (19)

c.-23G>C Us 5´UTR 46 (9) 22(5) 44 (9)

c.296-75_-66dup 1 intron 4 0 0 1

c.363G>A 1 exon 5 4 4* 0

c.829A>G (Thr276Ala) 1 exon 8 1 1 1

c.843+11C>T 1 intron 8 5 5* 1

c.1169+42C>G intron 10 43 (12) 27(6) 47 (14)

c.1170-6C>T 1 intron 10 2 1 2

c.1339_1344dup 1 exon 12 2 0 2

c.1362+38G>C 3´UTR 34 (11) 22(3) 40 (14)

c.1362+518C>T 1 Ds 3´UTR 0 0 1

c.1362+523G>A 1 Ds 3´UTR 0 0 1

1 Novel variation

Us=upstream of, Ds=downstream of

* p < 0.05

Page 54: Genetic Studies on Recurrent Miscarriage

54

Two of the novel variations; c.363G>A in exon 5, and c.843+11C>T in intron 8 were

significantly more frequent in RM women compared to control women, indicating that

these variations may increase the risk of miscarriage. The c.363G>A variation was

detected in 4 RM women (8.7%) and in none of the control women (p=0.010). The

c.843+11C>T variation was detected in 5 RM women (10.9%) and in one of the control

women (1%) (p=0.014).

2.2 Thrombomodulin and Endothelial Protein C Receptor (II, unpublished)

Mutations in the TM or EPCR genes may cause thrombophilia in the mother, and thereby

constitute a risk factor for miscarriages. In addition, homozygous mutations in the foetus

may cause miscarriage by other mechanisms as well (Isermann et al., 2001). An important

role for the TM-protein C-EPCR system for placental development and maintenance of

pregnancy has been shown in mouse models (Healy et al., 1995; Gu et al., 2002). The

relevance of these mechanisms for pregnancy associated complications in humans has,

however, not yet been determined.

2.2.1 Variations in the TM gene

In this study we analysed the entire coding region of TM including exon-intron boundaries

and the 5´ and 3´ UTRs for mutations in 86 patients with RM and 191 controls. As a

result, one exonic and one 3´UTR sequence variation in the TM gene were detected (Table

4). These variations were detected in both patients and controls. There were no significant

differences in the allele or genotype frequencies between all RM patients or RM women

and controls for either of the sequence variations in the TM gene.

A novel 18 bp deletion in the 3´UTR of the TM gene, 1728+23_+40del, was detected

in five patients (6%) and five controls (3%). The variation is located 23 bp downstream of

the exon-intron boundary. The other variation detected in TM, c.1418C>T, is a previously

reported common polymorphism resulting in an amino acid substitution from an alanine to

a valine at codon 455. The variation is located in the sixth epidermal growth factor like

domain of the gene that is responsible for thrombin binding and thereby necessary for

protein C activation (Esmon, 1989) but has, however, previously been shown to be neutral

in respect to thrombophilia (van der Velden et al., 1991). The variation is likely to be

Page 55: Genetic Studies on Recurrent Miscarriage

55

tolerated according to the SIFT program, and was commonly detected in a homozygous

state in both patients and controls in our study.

Table 4. The location of the variations in the TM gene, predicted amino acid changes, and the numbers of

patients and controls carrying the rarer allele in a hetero/(homozygous) state. Numbering of the nucleotides

is relative to the adenine in the ATG start codon in the reference sequence (UCSC Genome Browser,

http://genome.ucsc.edu/).

TM sequence variations

RM women+

DNA variation male partners RM women controls

(predicted amino acid change) Location n=86 n=46 n=191

c.1418C>T (Ala455Val) exon 1 41 (5) 26(3) 84 (8)

c.1728+23_+40del1 3´UTR 5 3 5

1 Novel variation

2.2.2 Variations in the EPCR gene

As a result of screening 86 patients and 191 controls for mutations in EPCR two exonic

variations and two variations in the non-coding regions of the gene were detected (Table

5). The two non-coding substitutions detected in EPCR have been previously reported as

common polymorphisms. Variation c.323-20T>C is located 20 bp upstream of exon 3 and

variation c.717+16G>C is located in the 3´ UTR. The c.717+16G>C polymorphism has

been suggested to have a modifier effect on the risk of venous thromboembolism (VTE).

Individuals homozygous for the C allele have elevated levels of activated protein C and a

lower risk of VTE (Medina et al., 2004). In addition, a recent study suggests that the C

allele decreases the risk of RM in women with a factor V Leiden mutation (Hopmeier et

al., 2008).

Variation c.655A>G is located in exon 4. It is a non-synonymous change, predicting a

Ser219Gly change. The variation is predicted to be tolerated according to the SIFT

program and previous studies have concluded that this variation does not increase the risk

of VTE (Medina et al., 2004; Ireland et al., 2005). The 23 bp duplication, c.323-9_336dup

in EPCR exon 3, duplicates the preceding 23 bases and results in a premature STOP codon

downstream of the insertion point. The mutation leads to the formation of a truncated

receptor lacking the extracellular and transmembrane domains. The truncated protein does

not function properly because it is not localized on the cell surface, cannot be secreted

from the cells, and does not bind activated protein C (Biguzzi et al., 2001). We detected

Page 56: Genetic Studies on Recurrent Miscarriage

56

this variation in three samples; in one female patient, in one male partner, and in one

control woman. All four variations in EPCR were detected in both patients and controls.

There were no significant differences in the allele or genotype frequencies between all RM

patients or RM women and controls for any of the EPCR sequence variations. Based on

these results we cannot conclude a role for the detected EPCR variations in RM.

Table 5. The location of the variations in the EPCR gene, predicted amino acid changes, and the numbers of

patients and controls carrying the rarer allele in a hetero/(homozygous) state. Numbering of the nucleotides

is relative to the adenine in the ATG start codon in the reference sequence (UCSC Genome Browser,

http://genome.ucsc.edu/).

EPCR sequence variations

RM women+

DNA variation male partners RM women controls

(predicted amino acid change) Location n=86 n=46 n=191

c.323-20T>C intron 2 41 (15) 18(11) 95 (33)

c.323-9_336dup exon 3 2 1 1

(TyrProGlnPheLeuSTOP)

c.655A>G (Ser219Gly) exon 4 22 (1) 12 49 (2)

c.717+16G>C 3´UTR 41 (15) 19(11) 93 (34)

2.2.3 TM/EPCR variations and outcome of heparin treatment

Of the women included in this study, 42 had been treated with heparin and/or aspirin in

attempts to improve the pregnancy outcome. Many of the RM women in our study had a

full term pregnancy when treated with heparin, but not all (Ulander V-M, personal

communication). As mutations in the TM or EPCR genes may cause disturbances of the

uteroplacental vasculature, we investigated whether TM or EPCR variations could be

associated with pregnancy outcome after heparin/aspirin treatment. Unfortunately, the

number of patients in each group was too small to obtain statistically significant results

and therefore associations between a specific TM/EPCR variation and the outcome of the

heparin treatment could not be studied.

Page 57: Genetic Studies on Recurrent Miscarriage

57

2.3 p53 (III, unpublished)

p53 functions have proven important for normal development in mice. However, its role in

human embryonic development has not been studied enough to make conclusions about its

role at different developmental stages. As the level of p53 expression and the ability of

p53 to interact with other proteins can be disturbed by sequence variations within the

gene, we screened the p53 gene for variations in our series of RM patients.

2.3.1 Variations in the p53 gene

Nine different sequence variations were found in the p53 gene by screening 40 couples

and 6 women with RM and 191 controls using dHPLC (Table 6). Six variations were

confirmed by sequencing to be located in the non-coding regions, and three variations

were found within the coding regions of the gene. All the variations have been previously

reported as common polymorphisms (p53 knowledgebase; NCBI SNP database). Two of

the exonic variations; G12032A, A13399G are synonymous and do not cause a change of

an amino acid, while one of the exonic variations, G12139C, is a common non-

synonymous polymorphism changing the amino acid of codon 72 from an arginine to a

proline. These R72P variants have been shown to differ biologically and biochemically

(Thomas et al., 1999), and have differences in their ability to interact with basic elements

of the transcriptional machinery, to induce apoptosis and to suppress transformed cell

growth (Dumont et al., 2003; Pim and Banks, 2004). This variation has previously been

studied with respect to RM. Pietrowski et al. (2005) studied 175 RM women and found an

association between the R72P polymorphism and RM, while Coulam et al. (2006) studied

205 RM women and showed no significant differences in the frequencies of R72P

polymorphism between RM patients and controls. In this study, we did not detect any

significant association between the R72P polymorphism and RM.

We found the polymorphism C11992A, located 29 bp upstream of exon 4, to be

associated with RM. The C/A and A/A genotypes of the C11992A polymorphism were

shown to be significantly more frequent (p-value <0.05) in women with RM than control

women. Women carrying the C/A or A/A genotype had a more than twofold increased risk

for miscarriage (p=0.0414, OR 2.083, CI 1.018-4.259). Statistical calculations using the

X2-test were performed following a dominant genotype model, where the C/C genotypes

Page 58: Genetic Studies on Recurrent Miscarriage

58

were compared against the C/A and A/A genotypes. A C/A or A/A genotype was detected

in 32.6% of the women with miscarriages and in 18.9% of the controls. When comparing

the allele frequencies of the C11992A variation, there was no significant difference

between the patient and control groups.

Table 6. The location of the variations in the p53 gene, the predicted amino acid change, and the number of

patients and controls carrying the rarer allele in a hetero/(homozygous) state. Numbering of the nucleotides

and codons is relative to the sequence presented by the p53 knowledgebase (http://p53.bii.a-star.edu.sg)

2005.

p53 sequence variations

RM women+

DNA variation male partners RM women controls

(predicted amino acid change) Location n=86 n=46 n=191

11827 C>G intron 2 23 (6) 12(4) 42 (9)

11992 C>A intron 3 22 (2) 13(2)* 29 (7)

12032 G>A exon 4 2 0 1

12139 G>C (Arg72Pro) exon 4 38 (8) 22(3) 77 (8)

13399 A>G exon 6 2 1 0

13964 G>C intron 6 0 0 2

14181 C>T intron 7 5 3 16

14201 T>G intron 7 5 3 16

14766 T>C intron 9 1 0 2

* p < 0.05

2.3.2 p53 expression in RM placenta

Previous studies have shown that complete absence of p53 in mice can result in lethality,

reduced fertility, and high susceptibility to the effects of environmental stress. In addition,

deviations in expression levels of p53 can have serious detrimental consequences (Sah et

al., 1995; Nicol et al., 1995; Godley et al., 1996). Futhermore, a previous study showed

higher expression levels of apoptosis related genes in chorionic villi from RM patients

compared to controls (Choi et al., 2003). To further study the role of p53 in human

development, we studied the expression of p53 in placental tissue samples from

miscarriages of RM patients.

Placental tissue samples were available from 19 miscarriages from seven women.

These were subjected to immunohistological staining to study the expression of p53 in the

placenta. No abnormal expression patterns were detected in these samples (Figure 6). All

Page 59: Genetic Studies on Recurrent Miscarriage

59

samples showed heterogenic and primarily weak p53 expression. Stronger expression was

mainly seen in the cytotrophoblasts and syncytiothrophoblasts.

Ten of the placental tissue samples were obtained from couples who both had a C/C

genotype at position 11992, and 9 of the samples were from couples where one partner

was a C/A heterozygous and the other a C/C homozygote. According to parental

genotypes none of the pregnancies from which placental tissue were obtained were A/A

homozygotes and hence we were not able to study whether the 11992A/A genotype would

affect the expression of p53 in the placenta.

Figure 6. A representative p53 staining of placental tissue obtained from a miscarriage.The expression

patterns were similar in all 19 tissue samples. Stronger heterogeneous p53 expression is seen in the

cytotrophoblasts and syncytiotrophoblasts.

2.4 Mitochondrial genes (IV)

Recent studies suggest that mitochondrial functions in the oocyte may have a critical role

in human embryonic development. Dysfunctions that affect the capacity of the

mitochondria to produce ATP or regulate the apoptosis cascade have been suggested to

cause early demise of the embryo in humans (Van Blerkom et al., 1998; Van Blerkom,

2004). Accordingly, we hypothesized that some women with RM could be healthy carriers

of low-level heteroplasmic mtDNA mutations which would predispose to RM.

Page 60: Genetic Studies on Recurrent Miscarriage

60

2.4.1 Variations in the mitochondrial genome

In this study we screened 48 RM women and 48 age matched control women for

heteroplasmic mtDNA variations using dHPLC. The method is based on detection of

heteroduplexes in a sample and can therefore be used to detect heteroplasmic sequence

variations. However, by sequencing a number of homoplasmic deviations from the

reference sequence were also detected. Because all samples were not sequenced, we could

not assess the frequency of homoplasmic changes.

Table 7. The location of the variations in the mtDNA, predicted amino acid changes, and the numbers of

RM women and controls carrying the variation in a heteroplasmic state. Numbering of the nucleotides are

based on the reference sequence in the Mitomap database.

mtDNA sequence variations

DNA variation RM control

(predicted amino acid change) Location women women

16362 T>C non-coding 1 0

16519 T>C non-coding 1 1

73 A>G non-coding 1 1

146 T>C non-coding 1 1

152 T>C non-coding 1 1

195 T>C non-coding 1 0

573 ins +1/2 bp non-coding 1 0

4232 T>C (Ile309Thr) ND1 1 0

4703 T>C ND2 1 0

4722 A>G (Thr85Ala) ND2 1 0

4727 A>G ND2 3 1

4745 A>G ND2 1 2

8265 T>C (Leu27Pro) CO2 0 1

9758 T>C CO3 0 1

11404 A>G ND4 1 0

14653 C>T 1 ND6 1 0

14569 G>A ND6 1 0

14766 T>C (Ile7Thr) CYB 2 2

15257 G>A (Asp171Asn) CYB 0 1 1 Novel variation

The 19 different heteroplasmic variations detected are listed in Table 7. A

heteroplasmic mtDNA variation was detected in 13 RM women (27%) and 9 control

women (19%). Of the detected 19 variations, 18 have previously been reported as

polymorphic variations in the Mitomap or mtDB database, while the variation 14653C>T,

a synonymous change in the ND6 gene, has not been previously reported. Altogether, 12

different variations were detected within seven mitochondrial protein coding genes. No

Page 61: Genetic Studies on Recurrent Miscarriage

61

heteroplasmic variations were detected within the mitochondrial tRNA and mitochondrial

rRNA regions. Within the non-coding control region seven heteroplasmic variations were

detected. Seven of the variations located within the coding regions are synonymous

changes, while five of the changes are non-synonymous, predicted to change the amino

acid. A non-synonymous change was detected in four RM women and four control

women. The heteroplasmy levels of the non-synonymous variations were studied using

solid-phase minisequencing and the heteroplasmy levels are presented in Table 8.

To study if the 7 variations detected in the mtDNA control region affected the mtDNA

copy number, the mtDNA levels in samples with a variation in the control region (n=8)

were analyzed using quantitative real-time PCR (Q-PCR) and compared to the mtDNA

levels in samples with no heteroplasmic mtDNA variations (n=6). We did not detect

significant differences (p=0.3568) in the mtDNA copy number between samples with a

control region variation and control samples with no variations (Figure 7).

Table 8. Heteroplasmy levels of mtDNA variations.

Level of

Variation Sample heteroplasmy

195 T>C RM patient 85 %

placenta 50 %

4232 T>C (Ile>Thr) RM patient 10 %

4722 A>G (Thr>Ala) RM patient 30 %

8265 T>C (Leu>Pro) control 75 %

11404 A>G RM patient 10 %

placenta <5 %

placenta 15 %

placenta 15 %

14653 C>T RM patient 85 %

placenta 65 %

placenta >95%

placenta 70 %

placenta 40 %

14766 T>C (Ile>Thr) RM patient <5 %

RM patient <5 %

control <95 %

control <5 %

15257 G>A (Asp>Asn) control 25 %

Page 62: Genetic Studies on Recurrent Miscarriage

62

2.4.2 Analysis of placental samples

Analysis of human pedigrees segregating pathogenic mtDNA point mutations show large

differences in the proportion of mutant mtDNAs among siblings and between generations

as a heteroplasmic variation is transmitted from the mother to the offspring (Ghosh et al.,

1996; Blok et al., 1997). The rapid changes in mitochondrial DNA allele frequency

observed between generations have been explained by an mtDNA bottleneck in the germ

line. Because of this bottleneck, the mtDNA genotype in the embryo is largely determined

by the genotype of the few mitochondria in the primordial germ cell (Jenuth et al., 1996;

Shoubridge, 2000). Due to the mitochondrial genetic bottleneck new mtDNA mutations

are either lost during transmission or quickly reach very high levels.

Figure 7. MtDNA copy number analysis by Q-PCR in cases with mtDNA control region variations. Eight

lymphocyte samples and one placental sample with at least one variation in the mtDNA control region were

compared to control lymphocyte (n=6) and placental samples (n=6) without mtDNA variations.

To study the transmission of the mtDNA variations from the mother to the offspring

we studied the presence of the previously mentioned variations in placental samples.

Unfortunately, placental samples were available from only three women with a

heteroplasmic mtDNA variation; 195T>C (1 placental sample), 11404A>G (3 placental

samples), and 14653C>T (4 placental samples). By analyzing these 8 placental samples

using dHPLC, we were able to show that the variations were present in the placental

tissue. The heteroplasmy level of these variations was studied using solid-phase

minisequencing (Table 8). In three cases the proportion of the mtDNA variation was

increased in the placenta compared to the mother, and in the remaining cases the

placenta lymphocyte lymphocyte

control

placenta

control

MtDNA copy

number (%)

Page 63: Genetic Studies on Recurrent Miscarriage

63

proportion was decreased in the placenta. This result is similar to previous studies

showing variable levels of heteroplasmic variations in the offspring (Ghosh et al., 1996;

Blok et al., 1997).

In addition, the mtDNA level of one placental sample with a control region variation

was compared to the mtDNA levels in placental samples without a control region variation

(n=6) using Q-PCR (Figure 7). No significant differences in the mtDNA copy number

levels were detected.

3 The impact of polymorphisms on phenotype

Single nucleotide polymorphisms (SNPs) are the most common type of human genetic

variation and to date over 3 million SNPs have been reported (Balakrishnan et al., 2005).

Functional SNPs can be categorized as coding SNPs that change amino acids, and

regulatory SNPs that modulate expression or splicing of the gene. However, the role of a

specific polymorphism in a disease or a phenotype may be difficult to assess, especially if

the polymorphism is located in the non-coding regions of the genome. Most

polymorphisms are neutral but a number of SNPs in both coding and non-coding regions

of the genome have also been reported to have phenotypic effects (Suh and Vijg, 2005;

Balakrishnan et al., 2005).

3.1 Variations in the nuclear genome (I, II, III, unpublished)

Most of the variations detected in the candidate genes AMN, TM, EPCR, and p53 were

more frequent than 1% in the control population, which is one of the definitions of a

genetic polymorphism. In recessive disorders it is, however, not uncommon to have carrier

frequencies of known deleterious mutations >1%, particularly in certain populations.

Considering the frequency of miscarriage in the general population, one would expect a

fairly high carrier frequency, if there was a relatively common mutation/variation

underlying miscarriages. Therefore, the frequency of a possible mutation alone is not

always a good indicator of whether a variation is a polymorphism or a mutation. It is also

difficult to determine the function of a SNP on the basis of the nucleotide sequence alone.

This is particularly true when the SNP neither alters an amino acid nor disrupts a well-

characterized motif that affects protein function or structure (Strahan and Read, 2004).

Page 64: Genetic Studies on Recurrent Miscarriage

64

In general, nucleotide substitutions within exons may influence the function of the

protein through changes in amino acid composition. Exonic variations that do not result in

a change to the amino acid sequence are called silent mutations, and their phenotypic

effect is often unclear. However, variations in the non-coding regions may also disturb or

alter protein function. SNPs in the promoter region of a gene may influence the

transcriptional activity of the gene by modulating transcription factor binding and

ultimately levels of the gene product. Polymorphisms in the noncoding sequences of

introns and the 3' and 5' UTRs may also result in defects in mRNA processing or stability

(Rebbeck et al., 2004; Malcolm, 2005; Suh and Vijg, 2005). Recently, a growing number

of polymorphisms within the non-coding regions of genes have been shown to alter the

expression or the splicing of a protein (Varley et al., 1998; Attanasio et al., 2001; Wang

XB et al., 2002; Wan et al., 2005; Bourdon et al., 2005). Therefore, some of the

polymorphisms detected in this study may disturb the normal functions of the proteins

encoded by the genes and thereby contribute to the etiology of miscarriage. It would be

interesting to further study the role of some of the variations in RM, especially the

deletion in TM, the duplication in EPCR, and the novel variations c.363G>A and

c.843+11C>T in AMN shown to be more frequent in RM women.

Although the majority of p53 mutations are missense mutations, generally causing a

loss in the ability to bind DNA, intronic variations may also alter the function of the

protein, or the binding sites for interacting proteins (Cho et al., 1994). Single base pair

substitutions in intron 4 of the p53 gene have been shown to disturb the binding of

unidentified transcription factors, resulting in decreased expression of mouse p53

(Beenken et al., 1991). This indicates that the p53 11992 variation may also affect protein

function. In humans, functional analysis has demonstrated elevated functional activity of

variation G13964C in intron 6, which further indicates the importance of intronic

sequences in regulating p53 function (Lehman et al., 2000). In addition, other intronic

base substitutions and duplications/deletions in the p53 gene sequence have been

associated with an increased risk for cancer, indicating that these variations may alter

splicing, expression levels or binding sites, and thereby the function of the protein

(Voglino et al., 1997; Wang-Gohrke et al., 1999; Lacerada et al., 2005). The mechanism

by which the introns regulate either transcription or mRNA stability is, however, unknown

but these results suggest that also the intronic C11992A polymorphism, which we detected

more frequently in patients than controls, may affect the expression level of p53.

Page 65: Genetic Studies on Recurrent Miscarriage

65

Alternatively, the variation may affect the protein functions required for protecting the

embryo from teratogens and environmental stress. Further studies are, however, necessary

to the define whether the intronic polymorphism has functional consequences and to

determine if the association with a higher risk for miscarriage is clinically significant.

3.1.1 Possible splice site variations

Approximately 60% of the listed disease causing mutations are missense or nonsense

mutations. In addition, mutations at splice sites make a significant contribution to human

genetic diseases. Approximately 15% of disease causing mutations affect pre-mRNA

splicing through exon skipping, activation of cryptic splice sites, creation of pseudo exons

within introns, or intron retention (Krawczak et al., 1992; Nakai and Sakamoto, 1994;

Baralle and Baralle, 2008). Therefore, we analysed the variations which were not

previously defined as common polymorphisms and the novel sequence variations with the

ESEfinder to see if any of these variations may affect splicing by altering the binding sites

of Ser/Arg-rich (SR) proteins. The program predicted that three single nucleotide

substitutions detected in the AMN gene and the deletion detected in the TM gene may

affect splicing by deleting at least one binding site for known SR-proteins and may

therefore result in exon skipping or intron retention (Table 9 and Figure 8). These

variations may affect splicing of the gene and thereby its function. Conformation of the

phenotypic effect of these variations, however, requires further studies.

The ESEfinder was mainly developed for analysing exonic sequences, but has also

been used to analyze and identify mutations in introns that affect splicing (Gabut et al.,

2005; Stogmann et al., 2006). However, the program only makes predictions about the

effect a variation may have on mRNA splicing. Thus, the presence of a high score motif in

a sequence does not necessarily mean that the sequence determines a splicing enhancer. In

addition, the sequences with the maximum scores are not necessarily the most effective

splicing enhancers. The program currently searches for ESE motifs corresponding to four

SR-proteins (SF2/ASF, SC35, SRp40, and SRp55) but there are several other SR-proteins

for which the ESE motifs have not yet been identified (Cartegni et al., 2003).

Page 66: Genetic Studies on Recurrent Miscarriage

66

Table 9. Threshold (Thr) values defined by the ESEfinder and predictive values obtained for normal and

variant alleles for sequence variations predicted to affect the binding sites of SR-proteins.

Figure 8. Diagram of the predicted values obtained by the ESEfinder for the exonic variation AMN

c.829A>G. Graphics of the normal (A) and variant (B) sequence show that the binding site for SF2/ASF

(black bar) present in the normal sequence (GACACCT) is abolished in the variant sequence (GACGCCT).

3.1.2 Couple analysis

The 40 couples included in this study were analysed to determine if some of the possible

functional variations could be detected in both partners of a couple, thereby enabling a

homozygous foetus. In two of the couples both partners had a different exonic or putative

splice site affecting variation in the AMN gene. One couple had AMN variations c.363G>A

and c.1170-6C>T, and another couple had AMN variations c.843+11C>T and

c.1339_1344dup. In these couples there is a 25% chance for every conceptus to be a

compound heterozygote for potentially deleterious AMN mutations in the two alleles.

Additionally, the newly identified 1728+23_+40 deletion in TM was detected in a

heterozygous state in both partners of one couple, enabling homozygosity of the deletion

SF2/ASF SC35 SRp55

Thr =1.956 Thr =2.383 Thr =2.267

AMN c.829A>G

normal allele 2.080 (GACACCT)

variant allele <1.956 (GACGCCT)

AMN c.843+11C>T

normal allele 2.506 (CGCGCGG) 2.584 (GGGCCGCG)

variant allele <1.956 (TGCGCGG) <2.383 (GGGCTGCG)

AMN c.1170-6C>T

normal allele 2.633 (GCCCTCAG) 2.922 (TACGCC)

variant allele <2.383 (GCTCTCAG) <2.267 (TACGCT)

TM c.1728+23_+40del

normal allele 2.742 (GGAGCCTG) 3.831 (TCCGTC)

variant allele sequence deleted sequence deleted

A) B)

Page 67: Genetic Studies on Recurrent Miscarriage

67

in the conceptus. This variation, located 23 bp downstream of the exon-intron boundary,

may affect splicing, or alternatively may have an effect on the stability of the mRNA.

Unfortunately, there were no tissue samples from miscarriages experienced by these

couples to study whether these variations were inherited in a homozygous form.

3.2 Variations in the mitochondrial genome (IV)

Although mtDNA is highly polymorphic and both pathogenic and silent variations are

generated in the mtDNA genome, the occurrance of heteroplasmic mtDNA sequence

variations is rare (Jenuth et al., 1996; McKenzie et al., 2004). For example, MELAS and

MERRF mutations are invariably heteroplasmic and homoplasmic mutations in these

genes are expected to be lethal (Howell et al., 2000). As the presence of two different

mtDNA populations in an individual is reported to be rare (Jenuth et al., 1996; DiMauro

and Schon, 2001; McKenzie et al., 2004), the percentage of samples with a heteroplasmic

variation was unexpectedly high in our study. According to our study every fifth control

woman carries a heteroplasmic variation, indicating that heteroplasmy is more common

than previously reported. However, it may be that the high level of homoplasmy is over-

emphasized as most studies rely on techniques that cannot detect low-level heteroplasmic

mtDNA variations (Grzybowski et al., 2000). The fact that we detected heteroplasmy in so

many of the studied samples is most likely due to the sensitivity of the detection method

used in our study. In 8/19 samples (42%) in which the level of heteroplasmy was tested

the ratio of variant sequences was ≤10% or ≥90%, and could not be detected by direct

sequencing.

In this study, a heteroplasmic variation, predicted to change an amino acid, was

detected in four RM women. All of these non-synonymous variations are previously

reported as homoplasmic polymorphisms, and are therefore likely to be non-pathogenic. It

is, however, possible that these variations are neutral when present in a small percentage

of cells, but pathogenic when a threshold is reached or in a homoplasmic state. In general,

the proportion of mutant mtDNAs in a heteroplasmic individual is expected to be very

high before a biochemical or clinical phenotype is observed (Jenuth et al., 1996). Non-

synonymous changes in mitochondrial genes may affect ATP production or apoptosis,

thereby leading to complications during early development. In addition, the mitochondria

are involved in the placental progesterone synthesis required for the maintenance of

Page 68: Genetic Studies on Recurrent Miscarriage

68

pregnancy (Tuckey, 2005). One of the detected coding region variations is novel, not

previously reported in the MitoMap or the mtDB database. This variation, 14653C>T in

the ND6 gene is, however, unlikely to be pathogenic as it does not change an amino acid.

In addition, this variation was detected at a heteroplasmy level of 85% in one otherwise

healthy RM woman. Consequently, this novel variation is likely to be neutral.

Analysis of the mtDNA control region revealed 7 heteroplasmic mutations among RM

patients. These variations in the most polymorphic region of the mitochondrial genome

have previously been reported to be polymorphisms. However, mutations in the control

region may affect replication of the mitochondrial genome and such mutations are

suggested to be an important factor in the outcome of a pregnancy. As mitochondrial

replication begins after implantation, replication defects are not likely to compromise

preimplantation embryogenesis, but depending upon mutant load, may lead to post-

implantation

embryonic death (Van Blerkom, 2004). It has been reported that the

localization of mitochondria in the early embryo is strictly regulated and the mitochondrial

distribution is thought to play a role in defining embryonic patterning and long term

viability of the blastomere. Therefore, mtDNA copy number may be an important

determinant of oocyte quality as too few copies would result in maldistribution in the early

embryo (Van Blerkom et al., 2000; Dumollard et al., 2007). To study if the detected

control region variations is affecting mtDNA replication, we analyzed the mtDNA copy

number in these samples. No significant differences in the mtDNA copy number were

shown.

4 Analysis of sex chromosome characteristics (V)

Recent publications have reported associations between RM and skewed XCI in females

and Y chromosome microdeletions in males. Therefore, both of these sex chromosome

characteristics were studied in our patients.

4.1 X chromosome inactivation analysis

In recent years several studies have shown that skewed XCI, defined as preferential

inactivation of one of the two X chromosomes, is increased in women with RM (Lanasa et

al., 1999; Lanasa et al., 2001; Uehara et al., 2001; Kuo et al., 2008). Normally inactivation

Page 69: Genetic Studies on Recurrent Miscarriage

69

of the X chromosomes occurs randomly on both the maternal and paternal X

chromosomes during the late blastocyst state of embryogenesis (Lyon, 1961). The

suggested explanation for the association between skewed XCI and RM is an unknown X

chromosomal mutation resulting in both preferential inactivation of the mutated X and

lethality in male embryos carrying the mutation (Lanasa et al., 1999; Uehara et al., 2001;

Bagislar et al., 2006).

To study the role of XCI in RM in our Finnish patients the XCI patterns were

examined in 46 women with RM and 95 control women. The XCI ratio was determined by

analysing the methylation status of the X-linked AR gene using a methylation sensitive

restriction enzyme that digests only nonmethylated DNA segments (Figure 9). Samples

homozygous at the AR (CAG)n locus were excluded from the X-inactivation study due to

the inability to distinguish between the two alleles. 3/46 women with RM and 14/95

control women were homozygous for the AR (CAG)n polymorphism and excluded from

further analysis. The ratio of inactivation between the two X chromosomes was calculated

for the remaining 43 female patients and 81 controls.

Figure 9. Fluorescence based X-inactivation analysis using a PCR based AR methylation assay. The

methylation sensitive restriction enzyme HpaII digests and prevents amplification of nonmethylated (active

X) DNA segments. (a) An example of equivalent inactivation of both alleles. (b) An example of extremely

skewed inactivation of the alleles. Undigested DNA is used as control for peak intensities of the alleles.

There are several possible reasons for the discrepancy between this study and those

previously reported. Firstly, it is possible that the number of patients is too small to

provide statistically reliable results if skewed XCI is the cause of RM in only a small

subset of patients. However, previous studies have included approximately the same

number of subjects and have been able to show an association between skewed XCI and

undigested undigested

Hpa II digested Hpa II digested

a) b)

Page 70: Genetic Studies on Recurrent Miscarriage

70

RM (Lanasa et al., 1999; Uehara et al., 2001). Secondly, previous studies have used

different inclusion criteria, which may affect the results. Our study included

patients with

three or more pregnancy losses. In contrast, other studies have included patients with only

two pregnancy losses (Lanasa et al., 1999; Lanasa et al., 2001; Kim et al., 2004). Finally,

in our study the patients and controls were of same age, the mean age in both groups being

31 years, whereas in other studies the patients were significantly older than their

controls

(Lanasa et al., 1999, Lanasa et al., 2001). Because skewed XCI increases in peripheral

lymphocytes with age due to depletion of the haematopoietic stem cell pool the age

discrepancy in other studies could have resulted in a bias towards an increase in skewed

XCI in RM cases when compared with younger controls (Sharp et al., 2000; Hogge et al.,

2007).

There are, however, also studies that have not detected an association between skewed

XCI and RM (Sullivan et al., 2003; Kim et al., 2004; Hogge et al., 2007; Dasoula et al.,

2008). Beever and coworkers (2003) evaluated the XCI status in over 200 women with

RM and did not detect an association between skewed XCI and RM. Futhermore, they

detected a significant excess of boys in live births among women with RM, which is

contrary to expectations if the skewed XCI is due to X-linked lethal mutations (Beever et

al., 2003). In some women skewed XCI may be due to the limited number of precursor

cells present during the late blastocyst stage when the X chromosome inactivation process

begins. At this developmental stage, there are approximately only 10–20 cells that will

subsequently form the embryo; the other cells differentiate into extra embryonic tissues.

Thus, some women may show skewed XCI due to chance alone (Lyon, 1971; Belmont,

1996; Gale et al., 1997). Alternatively, it is possible that skewed XCI is a heritable genetic

trait that is detected in family members in multiple generations. Genetic loci that influence

the selection of the inactivated X chromosome have been detected in mice (Percec et al.,

2002; Chadwick et al., 2006). Whether similar loci exist in humans

is not clear, but skewed

XCI has been mapped in humans to Xq28 (Pegoraro et al., 1997; Naumova et al., 1998). In

addition, skewed XCI appears to be inherited in some families and in these families

skewed XCI may be caused by a genetic variation that is not related to RM (Naumova et

al., 1996).

Page 71: Genetic Studies on Recurrent Miscarriage

71

4.2 Y chromosome microdeletion analysis

To date the knowledge of male contribution to RM is largely limited to karyotyping.

Karyotyping gives an overview of the whole genome, but reveals only large chromosomal

rearrangements. Submicroscopic duplications and deletions, such as Yq microdeletions,

remain unidentified using this method. Yq microdeletions are reported to cause

spermatogenetic failure and male infertility (Vogt et al., 1996; Hopps et al., 2003). There

are at least three regions, AZFa, AZFb, and AZFc, on the long arm of the Y chromosome,

in which deletions are associated with varying degree of spermatogenic failure. Although

it is widely accepted that deletions of these regions contribute to male infertility, the

function of the genes located in the AZF regions is poorly understood (Krausz and

McElreavey, 1999; Noordam and Repping, 2006; Ferlin et al., 2007). It has been reported

that men with Y chromosome microdeletions have low sperm counts and a higher

incidence of sperm chromosomal abnormalities (Foresta et al., 2005). Because sperm

chromosome abnormalities have been found in some men of RM couples (Rubio et al,

1999; Al-Hassan et al., 2005) Dewan and coworkers (2006) studied if Y chromosome

microdeletions are associated with RM. They reported that a significant portion of male

partners (82%) of women with RM had Y chromosomal microdeletions, and suggest that

the Y chromosomal genes may be important for early embryonic development in male

embryos (Dewan et al., 2006).

In our study we investigated if a similar association between Y chromosome

microdeletions and RM was present in our set of Finnish patients. We tested 40 male

partners of women with unexplained RM for the presence of Y chromosome

microdeletions. In addition, to the four STSs used in the previous study (Dewan et al.,

2006) we tested our patients using 33 additional STSs to enable the analysis of all AZF

regions. One STS was used to test for the presence of the short arm of the Y chromosome,

the other 32 STS loci were located on the long arm of the Y chromosome. Amplification

products for all of the 37 STSs used were detected in all tested patients, and thereby we

did not detect any Y chromosome microdeletions in our samples.

Although our study was performed on a bigger set of samples compared to the study

by Dewan et al. (2006), our study revealed no microdeletions and therefore does not

support the previously reported findings. It is possible that the involvement of

microdeletions in the aetiology of RM differs between populations and this may be the

Page 72: Genetic Studies on Recurrent Miscarriage

72

cause of different results between the studies. It is known that different populations have

different Y chromosome haplogroups and it has been shown that the susceptibility to Y

chromosomal microdeletions varies between different haplogroups (Arredi et al., 2007;

Yang et al., 2008). There may also be some technical problems affecting the results of the

Dewan et al. (2006) study (Noordam et al., 2006). These technical factors could be the

underlying cause of the discrepancy between the previously reported results and the results

of our study. According to the NCBI UniSTS Database, the STS marker DYS262 is

located on the short arm of the Y chromosome, and the markers DYS220, DYF86S1, and

DYF85S1 on the long arm in the AZFc region. Dewan et al. (2006) report that they found

patients with both DYS262 and an additional marker deleted. This indicates that these

patients would have multiple microdeletions or alternatively a region as big as 15 Mb

deleted. This is, however, unlikely in otherwise healthy males.

According to guidelines provided by the European Academy of Andrology (EAA) and

the European Molecular Genetics Quality Network (EMQN) for diagnostic laboratories,

the use of 6 STSs should be able to detect up to 95% of all reported Y microdeletions in

the AZF regions (Simoni et al., 2004). We tested our male patients using all the STSs

recommended in the guidelines, and in addition, over 30 other markers spanning the Y

chromosome. Therefore, it is unlikely that we would have missed at least any of the

known deletions.

Although we did not detect any microdeletions, it is important to keep in mind that a

normal lymphocyte karyotype does not exclude sperm abnormalities. High rates of

chromosomal abnormalities in sperm have been observed in RM patients with normal

lymphocyte karyotypes because the highly repetitive structure of the Y chromosome

sequence predisposes to de novo deletions (Rubio et al, 1999; Al-Hassan et al., 2005).

This study was performed using DNA extracted from peripheral blood and therefore, it is

possible that abnormalities leading to miscarriage have arisen in the spermatocytes during

spermatogenesis through de novo mutations.

Page 73: Genetic Studies on Recurrent Miscarriage

73

5 Limitations of the study

When considering the known causes of miscarriages, the aetiology of RM seems very

heterogeneous. Therefore, it is possible that mutations in the candidate genes of this study

cause miscarriage in only a small subpopulation of couples with miscarriages and one may

argue that more couples would be needed to determine the exact role of the candidate

genes in RM. After excluding the common variations, the mutation rate was low in the

candidate genes. This may be due to the fact that mutations in these genes are rare in

general or that they are a rare reason for RM. Screening more patients and controls may

also have revealed a statistical difference in the frequencies of the variations between

patients and controls. Additionally, a larger number of controls would be needed to

calculate the actual frequency of the novel variations and to determine their significance,

as when studying recessive conditions or complex diseases some of the controls are also

expected to be carriers of the disease causing mutation or susceptibility alleles.

Additionally, we also may have missed some mutations in the candidate genes.

Although in theory dHPLC is a highly sensitive and specific method for mutation

detection, the sensitivity of mutation detection by heteroduplex analysis is unlikely to be

100%. There are some rare instances where mutations may not be detected. Some

variations can only be detected at one unique temperature, and variations located within a

GC-rich sequence of a fragment with otherwise normal nucleotide ratios may in some

cases not be detected due to the high melt temperature of GC-rich pockets (Xiao and

Oefner, 2001). In this study one variation in TM (c.1418C>T) was initially not detected

using dHPLC but by sequencing a random selection of samples. Another potential

problem relates to the detectability of heteroplasmic mtDNA mutations in samples with a

proportion of <5% or >95% mutant mtDNA, because heteroduplex analysis becomes

unreliable beyond those limits (Biggin et al., 2004; Lim K et al., 2008).

Another limitation of this study is the use of blood samples for detection of

heteroplasmic mtDNA variations. It has been shown that the degree of heteroplasmy for

some mitochondrial DNA mutations varies considerably among tissues. Studies have

reported a decrease in the proportion of mutant mtDNA in blood by approximately 0.5%

to 2% per year (Howell et al., 2000; Rajasimha et al., 2008; Lim et al., 2008). As a

consequence blood may contain lower levels of the mutation compared to other tissues,

making the detection of low-level mutations demanding.

Page 74: Genetic Studies on Recurrent Miscarriage

74

Conclusions

The aim of this study was to find new genetic causes for RM, a pregnancy complication

that affects around 1% of couples trying to conceive. To date there are many factors

known to cause RM, but the underlying cause of a couple´s miscarriages often remains

unresolved as RM is a very heterogeneous condition. The identification of the underlying

causes of miscarriage is important for developing more successful treatments for women

experiencing RM and thereby preventing subsequent miscarriages. Identification of

underlying causes also helps to identify pregnant women that are at a risk of having a

miscarriage. This study aimed to identify new genetic factors underlying RM by studying

candidate genes - either identified from animal models or regulating functions important in

different stages of early foetal developmental - and sex chromosome characteristics

previously reported to be associated with RM.

By screening the couples for mutations in the AMN, TM and EPCR genes, two novel

AMN variations were shown to be associated with RM. One variation, c.363G>A, is an

exonic synonymous variation, and the other, c.843+11C>T is an intronic variation

predicted to affect splicing. These variations may increase the risk of RM in carrier

women. Further studies would, however, be required to confirm the phenotypic effect of

these variations. In addition, four interesting exonic or potential splice site disrupting

variations were detected in these genes; c.1339_1344dup and c.1170-6C>T in AMN,

c.1728+23_+40del in TM, and c.323-9_336dup in EPCR. In one couple both partners were

heterozygous for a deletion in the TM gene and in two couples both partners had different

exonic or potential splice site disrupting AMN variations. These variations may play a role

in the miscarriages of these couples. However, the phenotypic effects of the potentially

deleterious variations cannot be determined without further investigations. It would be

interesting to study these variatons in a larger sample set or alternatively, in

foetal/placental samples. However, most of the variations detected in these genes are

likely to be silent polymorphisms and, taken together, the results of this study suggest that

mutations in AMN, TM, and EPCR do not have a major role in RM in the patients studied,

even though their contribution to RM cannot be totally excluded.

Page 75: Genetic Studies on Recurrent Miscarriage

75

Additionally, a polymorphism in the p53 gene, C11992A, was shown to be associated

with RM in Finnish patients. The results indicate that women with a C/A or A/A genotype

have a twofold higher risk for RM than women with a C/C genotype. As the

polymorphism was, however, detected in a large number of controls as well it is likely to

be a contributing factor increasing the risk of RM in combination with other sequence

variations or environmental factors, rather than the sole cause of RM. It is also possible

that, instead of the 11992 locus, another genetic variation linked to this locus, is the actual

susceptibility locus increasing the risk for RM. While the exact role of p53 in development

is unclear, it is difficult to predict whether the variation disturbs functions needed for

normal embryonic development. The results of the p53 expression analysis indicate that

the miscarriages experienced by the studied couples are not due to altered p53 levels in the

placenta.

When screening the mitochondrial genome heteroplasmic mtDNA variations were

detected in 13 RM women, which was unexpected as heteroplasmic variations are reported

to be rare. Most of these variations are previously reported as polymorphisms and the

novel variation is likely to be non-pathogenic as it is synonymous. However, the role of

these variations in the aetiology of miscarriage cannot be excluded as some variations may

be pathogenic only when a certain threshold is reached. The transmission of the variations

was studied by analysing placental samples and the result is similar to previous studies

showing that a heteroplasmic variation in a mother is transmitted to her offspring in

different proportions. This supports the hypothesis that a woman with no disease

phenotype can carry a low-level heteroplasmic variation that may cause foetal death when

transmitted to the foetus in a higher proportion. However, the variations detected in this

study were judged to be polymorphisms, and not likely to contribute to the miscarriages

experienced by the women studied. This, however, does not exclude the role of other

heteroplasmic mtDNA variations in miscarriage.

In this study sex chromosome abnormalities, recently reported to be associated with

RM, were also investigated to study if these factors could underlie the RM in Finnish

patients. The female patients were studied for skewed X chromosome inactivation and the

male partners were studied for Y chromosome microdeletions. Based on the results of the

study it was concluded that there is no association between skewed X chromosome

inactivation or Y chromosome microdeletions and RM in Finnish patients.

Page 76: Genetic Studies on Recurrent Miscarriage

76

Although no major genetic cause explaining the miscarriages in the patients could be

identified, the results of the genealogic studies indicate that a subset of patients possibly

have a common genetic cause underlying their miscarriages. The results of the genealogic

studies suggest that there may be an eastern enrichment of ancestral birthplaces in the

province of Kuopio, which was unexpected as the patients were collected from Helsinki. It

would be interesting to collect more patients from this region for further genetic studies.

Page 77: Genetic Studies on Recurrent Miscarriage

77

Acknowledgements

This study was carried out at the Folkhälsan Institute of Genetics, and the Department of

Medical Genetics, University of Helsinki during 2004-2009. I wish to thank the former

and current heads of these institutes, Professors Anna-Elina Lehesjoki, Kristiina Aittomäki

and Päivi Peltomäki for providing excellent working facilities in Biomedicum Helsinki.

I am grateful to the following organizations for financially supporting my work: the

Folkhälsan Research Foundation, the Sigrid Juselius Foundation, the Biomedicum

Helsinki Foundation, the Emil Aaltonen Foundation, the Maud Kuistila Memorial

Foudation, the University of Helsinki Funds, the Oskar Öflund Foundation, and Svenska

Kulturfonden.

I wish to express my gratitude to my supervisor docent Kristiina Aittomäki for giving me

the opportunity to work in her research group and letting me participate in a very

interesting project. I am grateful to Kristiina for giving me responsibility for my projects

and for her confidence in me throughout this project. I guess you knew I would end up

defending my thesis eventhough I wasn´t so convinced myself all the time. I would also

like to thank Kristiina for supporting me in my aim of becoming a medical geneticist.

Docent Leena Anttila and docent Nina Horelli-Kuitunen, the official reviewers of this

thesis work, are warmly thanked for critical review and for valuable suggestions that

improved this thesis.

Professor Päivi Peltomäki is warmly thanked for accepting the important post of custos at

the dissertation.

I do owe enormous thanks to Jodie Painter, my supervisor in the lab, for encouragement

and for always being willing to help me. Jodie is thanked for help in the lab, especially in

the beginning of the project, invaluable help during a number of writing processes and

teaching me a scientific way of thinking. Jodie is also thanked for revising the language of

this thesis work and helpful comments to improve it.

I am deeply grateful to all the co-authors and collaborators in this study for their valuable

contributions to this thesis work. Veli-Matti Ulander and Risto Kaaja are thanked for

collecting and treating the patients included in this study, for clinical expertise and for a

fruitful and nice collaboration. Ralf Bützow is thanked for providing his expertise in

immunohistochemical studies. Anu Wartiovaara (née Suomalainen) and Alexandra Götz

are thanked for collaboration within the mitochondrial field and valuable contributions to

my thesis work under a tight schedule. Sarah Ariansen is warmly thanked for valuable

collaboration and for welcoming me to visit their lab in Oslo. Taru Ahvenainen is thanked

for careful lab work with the TM gene.

I am thankful to research nurses Katja Kuosa and Eija Kortelainen for providing me with

exact information about the study subjects. Aila Riikonen, Jaana Welin-Haapamäki,

Stephan Keskinen, Marjatta Valkama, and Solveig Halonen are acknowledged for helping

with practical matters during the years.

Page 78: Genetic Studies on Recurrent Miscarriage

78

I would like to thank Per-Henrik Groop for encouragement and interest in my studies,

thesis project, and future plans. Hannele Laivuori is thanked for fruitful collaboration

within the field of reproductive medicine and a positive attitude towards my work.

During my time working in the lab and writing my thesis I have greatly enjoyed the

company of my research colleagues at Folkhälsan Institute of Genetics. Maria, my fellow

PhD student in our small group, your company and support have been really important for

me, thank you. Vilma is thanked for friendship, interesting converations and for sharing

the frustration of working with the WAVE. I would like to thank Anne S, Eija, and Nina

for friendship and helping me with all kinds of questions, especially when finishing this

thesis project. Kati and Ann-Liz are thanked for their positive attitude towards most things

in life, and for making me smile :) Peter and Carol, the FHRM team, thank you for

amusing planning meetings during the past years. Saara, Mervi, Jaakko, Otto, Maria K,

Hanna V, Elina, Reetta, and all the former and present members of the Folkhälsan team

are thanked for useful advice, friendship, and for creating a nice and unique working

atmosphere.

I also wish to thank Joni, Annu, Helena, Pia, and Jeanette, with whom I have been sharing

the office room, for nice conversations about traveling and renovations, and for patience in

listening to and sharing my frustration when things were not going the way I wished. It

became quite empty in our room since you left.

I want to express my warmest thanks to my dear friends, Minttu, Ammi, and Stina for

always being there when I have needed you. Minja, Jenny, Anna, Mariann, Emma, and

Julia are warmly thanked for friendship, for scientific and non-scientific discussions

during both lunch hours and spare-time, and for making studing biology fun! It has been a

great pleasure to get to know you. I wish to sincerely thank Juhis and Toffe for friendship

and for convincing me since ages ago that I should do a PhD.

My deepest appreciation goes to my family for supporting me in all my decisions and an

encouraging attitude towards my studies. I own my warmest thanks to Miro, my friend,

my love, for his love and support during these years and for taking my thoughts off work.

Thank you for taking care of the penguin.

Finally, all the patients who have voluntarily participated in this study, thereby making it

possible, are warmly thanked.

Helsinki, February 2009

Page 79: Genetic Studies on Recurrent Miscarriage

79

References

Aarli A, Kristoffersen EK, Jensen TS, Ulvestad E, Matre R (1997). Suppressive effect on lymphoproliferation in vitro by soluble annexin II released from isolated placental membranes. Am J Reprod Immunol. 38:313-319.

Abdalla HI, Burton G, Kirkland A, Johnson MR, Leonard T, Brooks AA, Studd JW (1993). Age, pregnancy amd miscarriage: uterine versus ovarian factors. Hum Reprod. 8:1512-1517.

Aittomäki K, Lucena JL, Pakarinen P, Sistonen P, Tapanainen J, Gromoll J, Kaskikari R, Sankila EM, Lehväslaiho H, Engel AR, Nieschlag E, Huhtaniemi I, de la Chapelle A (1995). Mutation in the follicle-stimulating hormone receptor gene causes hereditary hypergonadotropic ovarian failure. Cell. 82:959-968.

Aittomäki K, Herva R, Stenman UH, Juntunen K, Ylöstalo P, Hovatta O, de la Chapelle A (1996). Clinical features of primary ovarian failure caused by a point mutation in the follicle-stimulating hormone receptor gene. J Clin Endocrinol Metabol. 81:3722-3726.

Aldrich CL, Stephenson MD, Karrison T, Odem RR, Branch DW, Scott JR, Schreiber JR, Ober C (2001). HLA-G genotypes and pregnancy outcome in couples with unexplained recurrent miscarriage. Mol Hum Reprod. 7:1167-1172.

Al-Hassan S, Hellani A, Al-Shahrani A, Al-Deery M, Jaroudi K, Coskun S (2005). Sperm chromosomal abnormalities in patients with unexplained recurrent abortions. Arch Androl. 51:69-76.

Aminoff M, Carter JE, Chadwick RB, Johnson C, Gräsbeck R, Abdelaal MA, Broch H, Jenner LB, Verroust PJ, Moestrup SK, de la Chapelle A, Krahe R (1999). Mutations in CUBN, encoding the intrinsic factor-vitamin B12 receptor, cubilin, cause hereditary megaloblastic anaemia 1. Nat Genet. 21:309-313.

Antonarakis SE, Beckmann JS (2006). Mendelian disorders deserve more attention. Nat Rev Genet. 7:277-282.

Armstrong JF, Kaufman MH, Harrison DJ, Clarke AR (1995). High-frequency developmental abnormalities in p53-deficient mice. Curr Biol. 5:931-936.

Arredi B, Ferlin A, Speltra E, Bedin C, Zuccarello D, Ganz F, Marchina E, Stuppia L, Krausz C, Foresta C (2007). Y-chromosome haplogroups and susceptibility to azoospermia factor c microdeletion in an Italian population. J Med Genet. 44:205-208.

Arredondo F, Noble LS (2006). Endocrinology of recurrent pregnancy loss. Semin Reprod Med. 24:33-39.

Ashley MV, Laipis PJ, Hauswirth WW (1989). Rapid segregation of heteroplasmic bovine mitochondria. Nucleic Acids Res. 17:7325-7331.

Attanasio C, de Moerloose P, Antonarakis SE, Morris MA, Neerman-Arbez M (2001). Activation of multiple cryptic donor splice sites by the common congenital afibrinogenemia mutation, FGA IVS4 + 1 G->T. Blood. 97:1879-1881.

Azofeifa J, Voit T, Hubner C, Cremer M (1995). X-chromosome methylation in manifesting and healthy carriers of dystrophinopathies: concordance of activation ratios among first degree female relatives and skewed inactivation as cause of the affected phenotypes. Hum Genet. 96:167-176.

Baek KH (2004). Aberrant gene expression associated with recurrent pregnancy loss. Mol Hum Reprod. 10:291-297.

Baek KH, Choi BC, Lee J-H, Choi H-K, Lee S-H, Kim J-W, Hill JA, Chung H-M, Ko JJ, Cha KY (2002) Comparison of gene expression at the feto-maternal interface between normal and recurrent pregnancy loss patients. Reprod Fertil Dev. 14:235–240.

Bagislar S, Ustuner I, Cengiz B, Soylemez F, Akyerli CB, Ceylaner S, Ceylaner G, Acar A, Ozcelik T (2006). Extremely skewed X-chromosome inactivation patterns in women with recurrent spontaneous abortion. Aust N Z J Obstet Gynaecol. 46:384-387.

Page 80: Genetic Studies on Recurrent Miscarriage

80

Bajekal N, Li TC (2000). Fibroids, infertility and pregnancy wastage. Hum Reprod Update. 6:614-620.

Balakrishnan VS, Rao M, Jaber BL; DialGene Consortium (2005). Genomic medicine, gene polymorphisms, and human biological diversity. Semin Dial. 18:37-40.

Balint EE, Vousden KH (2001). Activation and activities of the p53 tumour suppressor protein. Br J Cancer. 85:1813-1823.

Ballabio A (1993). The rise and fall of positional cloning? Nat Genet. 3:277-279.

Baralle D, Baralle M (2005). Splicing in action: assessing disease causing sequence changes. J Med Genet. 42:737-748.

Beddington RS, Robertson EJ (1999). Axis development and early asymmetry in mammals. Cell. 96:195-209.

Beenken SW, Karsenty G, Raycroft L, Lozano G (1991). An intron binding protein is required for transformation ability of p53. Nucleic Acids Res. 19:4747-4752.

Beever CL, Stephenson MD, Peñaherrera MS, Jiang RH, Kalousek DK, Hayden M, Field L, Brown CJ, Robinson WP (2003). Skewed X-chromosome inactivation is associated with trisomy in women ascertained on the basis of recurrent spontaneous abortion or chromosomally abnormal pregnancies. Am J Hum Genet. 72:399-407.

Belmont JW (1996). Genetic control of X inactivation and processes leading to X-inactivation skewing. Am J Hum Genet. 58:1101–1108.

Bergstrom CT, Pritchard J (1998). Germline bottlenecks and the evolutionary maintenance of mitochondrial genomes. Genetics. 149:2135-2146.

Biggin A, Henke R, Bennetts B, Thorburn DR, Christodoulou J (2005). Mutation screening of the mitochondrial genome using denaturing high-performance liquid chromatography. Mol Genet Metab. 84:61-74.

Biguzzi E, Merati G, Liaw PC, Bucciarelli P, Oganesyan N, Qu D, Gu JM, Fetiveau R, Esmon CT, Mannucci PM, Faioni EM (2001). A 23bp insertion in the endothelial protein C receptor (EPCR) gene impairs EPCR function. Thromb Haemost. 86:945-948.

Blok RB, Gook DA, Thorburn DR, Dahl HH (1997). Skewed segregation of the mtDNA nt 8993 (T-->G) mutation in human oocytes. Am. J. Hum. Genet. 60:1495-1501.

Bond GL, Hu W, Levine AJ (2005). MDM2 is a central node in the p53 pathway: 12 years and counting. Curr Cancer Drug Targets. 5:3-8.

Bose P, Black S, Kadyrov M, Weissenborn U, Neulen J, Regan L, Huppertz B (2005). Heparin and aspirin attenuate placental apoptosis in vitro: implications for early pregnancy failure. Am J Obstet Gynecol. 192:23-30.

Bourdon JC, Fernandes K, Murray-Zmijewski F, Liu G, Diot A, Xirodimas DP, Saville MK, Lane DP (2005). p53 isoforms can regulate p53 transcriptional activity. Genes Dev. 19:2122-2137.

Brown CJ, Robinson WP (2000). The causes and consequences of random and non-random X chromosome inactivation in humans. Clin Genet. 58:353-363.

Brown SD (1998). Mouse models of genetic disease: new approaches, new paradigms. J Inherit Metab Dis. 21:532-539.

Brown SD, Nolan PM (1998). Mouse mutagenesis-systematic studies of mammalian gene function. Hum Mol Genet. 7:1627-1633.

Bulletti C, Flamigni C, Giacomucci E (1996). Reproductive failure due to spontaneous abortion and recurrent miscarriage. Hum Reprod Update. 2:118-136.

Bussen S, Sutterlin M, Steck T (1999). Endocrine abnormalities during follicular phase in women with recurrent spontaneous abortion. Hum Reprod. 14:18-20.

Carrington B, Sacks G, Regan L (2005). Recurrent miscarriage: pathophysiology and outcome. Curr Opin Obstet Gynecol. 17:591-597.

Cartegni L, Wang J, Zhu Z, Zhang MQ, Krainer AR (2003). ESEfinder: A web resource to identify exonic splicing enhancers. Nuc Acid Research. 31:3568-3571.

Page 81: Genetic Studies on Recurrent Miscarriage

81

Carter AM (2007). Animal models of human placentation--a review. Placenta. 28(Suppl):S41-47.

Chadwick LH, Pertz LM, Broman KW, Bartolomei MS, Willard HF (2006). Genetic control of X chromosome inactivation in mice: definition of the Xce candidate interval. Genetics. 173:2103-2110.

Chavatte-Palmer P, Guillomot M (2007). Comparative implantation and placentation. Gynecol Obstet Invest. 64:166-174.

Cheah PL, Looi LM (2001). p53: an overview of over two decades of study. Malays J Pathol. 23:9-16.

Cho Y, Gorina S, Jeffrey PD, Pavletich NP (1994). Crystal structure of a p53 tumor suppressor-DNA complex: understanding tumorigenic mutations. Science. 265:346-355.

Choi H-K, Choi BC, Lee S-H, Kim J-W, Cha KY, Baek K-H (2003). Expression of angiogenesis- and apoptosis-related genes in chorionic villi derived from recurrent pregnancy patients. Mol Reprod. Dev 66:24–31.

Choi J, Donehower LA (1999). p53 in embryonic development: maintaining a fine balance. Cell Mol Life Sci. 55:38-47.

Clark DA, Coulam CB, Daya S, Chaouat G (2001). Unexplained sporadic and recurrent miscarriage in the new millennium: a critical analysis of immune mechanisms and treatments. Hum Reprod Update. 7:501-511.

Clifford K, Rai R, Watson H, Regan L (1994). An informative protocol for the investigation of recurrent miscarriage: preliminary experience of 500 consecutive cases. Hum Reprod. 9:1328-1332.

Cnattingius D, Signorello LB, Anneren G, Clausson B, Ekblom A, Ljunger E, Blot WJ, McLaughlin JK. Petersson G, Rane A, Granath F (2000). Caffein intake and the risk of first-trimester spontaneous abortion. N Eng J Med. 343:1839-1845.

Coan PM, Burton GJ, Ferguson-Smith AC (2005) Imprinted genes in the placenta--a review. Placenta. 26 Suppl A:S10-20.

Collins FS (1992). Positional cloning: let's not call it reverse anymore. Nat Genet. 1:3-6.

Collins FS (1995). Positional cloning moves from perditional to traditional. Nat Genet. 9:347-350.

Collins FS, Patrinos A, Jordan E, Chakravarti A, Gesteland R, Walters L (1998). New goals for the U.S. Human Genome Project: 1998-2003. Science. 282:682-689.

Costeas PA, Koumouli A, Giantsiou-Kyriakou A, Papaloizou A, Koumas L (2004). Th2/Th3 cytokine genotypes are associated with pregnancy loss. Hum Immunol. 65:135-141.

Coulam CB, Kay C, Jeyendran RS (2006). Role of p53 codon 72 polymorphism in recurrent pregnancy loss. Reprod Biomed Online. 12:378-382.

Coulam CB, Jeyendran RS (2008). Vascular endothelial growth factor gene polymorphisms and recurrent pregnancy loss. Am J Reprod Immunol. 59:301-305.

Couto E, Barini R, Zaccaria R, Annicchino-Bizzacchi JM, Passini Junior R, Pereira BG, Silva JC, Pinto e Silva JL (2005) Association of anticardiolipin antibody and C677T in methylenetetrahydrofolate reductase mutation in women with recurrent spontaneous abortions: a new path to thrombophilia? Sao Paulo Med J. 123:15-20.

Crawley JT, Gu JM, Ferrell G and Esmon CT (2002). Distribution of endothelial cell protein C/activated protein C receptor (EPCR) during mouse embryo development. Thromb Haemost. 88:259-266.

Cross JC (2006). Placental function in development and disease. Reprod Fertil Dev. 18:71-76.

Cross JC, Werb Z and Fisher SJ (1994). Implantation and the placenta: key pieces of the development puzzle. Science. 266:1508-1518.

Daher S, Shulzhenko N, Morgun A, Mattar R, Rampim GF, Camano L, DeLima MG (2003). Associations between cytokine gene polymorphisms and recurrent pregnancy loss. J Reprod Immunol. 58:69-77.

Page 82: Genetic Studies on Recurrent Miscarriage

82

Dahlbäck B, Villoutreix BO (2005). The anticoagulant protein C pathway. FEBS Lett. 579:3310-3316.

Dalton CF, Laird SM, Estdale SE, Saravelos HG, Li TC (1998). Endometrial protein PP14 and CA-125 in recurrent miscarriage patients; correlation with pregnancy outcome. Hum Reprod. 13:3197–3202.

Dasoula A, Kalantaridou S, Sotiriadis A, Pavlou M, Georgiou I, Paraskevaidis E, Makrigiannakis A, Syrrou M (2008). Skewed X-chromosome inactivation in Greek women with idiopathic recurrent miscarriage. Fetal Diagn Ther. 23:198-203.

den Dunnen JT, White SJ (2006). MLPA and MAPH: sensitive detection of deletions and duplications. Curr Protoc Hum Genet. Chapter 7.14.

Devine DH, Whitman-Elia G, Best RG, Edwards JG (2000). Paternal paracentric inversion of chromosome 2: a possible association with recurrent pregnancy loss and infertility. J Assist Reprod Genet. 17:293–296.

Dewan S, Puscheck EE, Coulam CB, Wilcox AJ, Jeyendran RS (2006). Y-chromosome microdeletions and recurrent pregnancy loss. Fertil Steril. 85:441-445.

Dhont M (2003). Recurrent miscarriage. Curr Womens Health Rep. 3:361-366.

DiMauro S (2001). Lessons from mitochondrial DNA mutations. Semin. Cell Dev. Biol. 12:397-405.

DiMauro S (2007) Mitochondrial DNA medicine. Biosci Rep. 27:5-9.

DiMauro S, Schon EA (2001). Mitochondrial DNA mutations in human disease. Am J Med Genet. 106:18-26.

Dittman WA, Majerus PW (1989). Sequence of a cDNA for mouse thrombomodulin and comparison of the predicted mouse and human amino acid sequences. Nucleic Acids Res 17:802.

Dizon-Townson DS, Lu J, Morgan TK, Ward KJ (2000). Genetic expression by fetal chorionic villi during the first trimester of human gestation. Am J Obstet Gynecol. 183:706-711.

Doggen CJ, Kunz G, Rosendaal FR, Lane DA, Vos HL, Stubbs PJ, Manger Cats V, Ireland H (1998). A mutation in the thrombomodulin gene, 127G to A coding for Ala25Thr, and the risk of myocardial infarction in men. Thromb Haemost. 80:743-748.

Dominguez-Rojas V, de Juanes-Pardo JR, Astasio-Arbiza P, Ortega-Molina P, Gordillo-Florencio E (1994). Spontaneous abortion in a hospital population: are tobacco and coffee intake risk factors? Eur J Epidemiol. 10:665-668.

Dumollard R, Duchen M, Carroll J (2007). The role of mitochondrial function in the oocyte and embryo. Curr Top Dev Biol. 77:21-49.

Dumont P, Leu JI, Della Pietra AC 3rd, George DL, Murphy M (2003). The codon 72 polymorphic variants of p53 have markedly different apoptotic potential. Nat Genet. 33:357-365.

Dvash T, Ben-Yosef D, Eiges R (2006). Human embryonic stem cells as a powerful tool for studying human embryogenesis. Pediatr Res. 60:111-117.

Egozcue S, Blanco J, Vendrell JM, Garcia F, Veiga A, Aran B, Barri PN, Vidal F, Egozcue J (2000). Human male infertility: chromosomal anomalies. Meiotic disorders, abnormal spermatozoa and recurrent abortion. Hum Reprod Update. 6:93-105.

Empson M, Lassere M, Craig JC, Scott JR (2002). Recurrent pregnancy loss with antiphospholipid antibody: a systematic review of therapeutic trials. Obstet Gynecol. 99(1):135-144.

Esmon CT (1989). The roles of protein C and thrombomodulin in the regulation of blood coagulation. J Biol Chem. 264:4743-4746.

Esmon CT (2000). The endothelial cell protein C receptor. Thromb Haemost. 83:639-643.

Faioni EM, Franchi F, Castaman G, Biguzzi E, Rodeghiero F (2002). Mutations in the thrombomodulin gene are rare in patients with severe thrombophilia. Br J Haematol. 118:595-599.

Page 83: Genetic Studies on Recurrent Miscarriage

83

Farquharson RG, Quenby S, Greaves M (2002). Antiphospholipid syndrome in pregnancy: a randomized, controlled trial of treatment. Obstet Gynecol. 100(3):408-413.

Fazel A, Vincenot A, Malassine A, Soncin F, Gaussem P, Alsat E, Evain-Brion D (1998). Increase in expression and activity of thrombomodulin in term human syncytiotrophoblast microvilli. Placenta. 19:261-268.

Ferlin A, Arredi B, Speltra E, Cazzadore C, Selice R, Garolla A, Lenzi A, Foresta C (2007). Molecular and clinical characterization of y chromosome microdeletions in infertile men: a 10-year experience in Italy. J Clin Endocrinol Metab. 92:762-770.

Foka ZJ, Lambropoulos AF, Saravelos H, Karas GB, Karavida A, Agorastos T, Zournatzi V, Makris PE, Bontis J, Kotsis A (2000) Factor V leiden and prothrombin G20210A mutations, but not methylenetetrahydrofolate reductase C677T, are associated with recurrent miscarriages. Hum Reprod. 15:458-462.

Foresta C, Garolla A, Bartoloni L, Bettella A, Ferlin A (2005). Genetic abnormalities among severely oligospermic men who are candidates for intracytoplasmic sperm injection. J Clin Endocrinol Metab. 90:152-156.

Franco B, Ballabio A (2006) X-inactivation and human disease: X-linked dominant male-lethal disorders. Curr Opin Genet Dev. 16:254-259.

Fryns JP, Van Buggenhout G (1998). Structural chromosome rearrangements in couples with recurrent fetal wastage. Eur J Obstet Gynecol Reprod Biol. 81:171-176.

Fyfe JC, Madsen M, Hojrup P, Christensen EI, Tanner SM, de la Chapelle A, He Q, Moestrup SK (2004). The functional cobalamin (vitamin B12)-intrinsic factor receptor is a novel complex of cubilin and amnionless. Blood. 103:1573-1579.

Gabut M, Mine M, Marsac C, Brivet M, Tazi J, Soret J (2005). The SR protein SC35 is responsible for aberrant splicing of the E1alpha pyruvate dehydrogenase mRNA in a case of mental retardation with lactic acidosis. Mol Cell Biol. 25:3286-3294.

Gale RE, Fielding AK, Harrison CN, Linch DC (1997). Acquired skewing of X-chromosome inactivation patterns in myeloid cells of the elderly suggests stochastic clonal loss with age. Br J Haematol. 98:512-519.

Ghosh SS, Fahy E, Bodis-Wollner I, Sherman J, Howell N (1996). Longitudinal study of a heteroplasmic 3460 Leber hereditary optic neuropathy family by multiplexed primer-extension analysis and nucleotide sequencing. Am. J. Hum. Genet. 58:325-334.

Girardi G (2005). Heparin treatment in pregnancy loss: Potential therapeutic benefits beyond anticoagulation. J Reprod Immunol. 66:45-51.

Godley LA, Kopp JB, Eckhaus M, Paglino JJ, Owens J, Varmus HE (1996). Wild-type p53 transgenic mice exhibit altered differentiation of the ureteric bud and possess small kidneys. Genes Dev. 10:836-850.

Gräsbeck R, Gordin R, Kanteroi I, Kuhlback B (1960). Selective vitamin B12 malabsorption and proteinuria in young people. Acta Med Scand. 167:289-296.

Grzybowski T (2000). Extremely high levels of human mitochondrial DNA heteroplasmy in single hair roots. Electrophoresis. 21:548-553.

Gu JM, Crawley JT, Ferrell G, Zhang F, Li W, Esmon NL, Esmon CT (2002). Disruption of the endothelial cell protein C receptor gene in mice causes placental thrombosis and early embryonic lethality. J Biol Chem. 277:43335-43343.

Haaf T (2006). Methylation dynamics in the early mammalian embryo: implications of genome reprogramming defects for development. Curr Top Microbiol Immunol. 310:13-22.

Healy AM, Rayburn HB, Rosenberg RD, Weiler H (1995). Absence of the blood-clotting regulator thrombomodulin causes embryonic lethality in mice before development of a functional cardiovascular system. Proc Natl Acad Sci USA. 92:850-854.

Healy AM, Hancock WW, Christie PD, Rayburn HB, Rosenberg RD (1998). Intravascular coagulation activation in a murine model of thrombomodulin deficiency: effects of lesion size, age, and hypoxia on fibrin deposition. Blood. 92:4188-4197.

Page 84: Genetic Studies on Recurrent Miscarriage

84

Hiatt BK, Lentz SR (2002). Prothrombotic states that predispose to stroke. Curr Treat Options Neurol. 4(6):417-425.

Hill JA, Choi BC (2000). Maternal immunological aspects of pregnancy success and failure. J Reprod Fertil. 55(Suppl):91–97.

Hogge WA, Prosen TL, Lanasa MC, Huber HA, Reeves MF (2007). Recurrent spontaneous abortion and skewed X-inactivation: is there an association? Am J Obstet Gynecol. 196:384.e1-8.

Hopmeier P, Puehringer H, van Trotsenburg M, Atamaniuk J, Oberkanins C, Dossenbach-Glaninger A (2008). Association of endothelial protein C receptor haplotypes, factor V Leiden and recurrent first trimester pregnancy loss. Clin Biochem. 41:1022-1024.

Hopps CV, Mielnik A, Goldstein M, Palermo GD, Rosenwaks Z, Schlegel PN (2003). Detection of sperm in men with Y chromosome microdeletions of the AZFa, AZFb and AZFc regions. Hum Reprod. 18:1660-1665.

Hoshimoto K, Komine F, Hayashi M, Ohkura T (2002). Plasma soluble Fas changes during early pregnancy and miscarriage. Clin Chim Acta. 323:157-160.

Howell N, Chinnery PF, Ghosh SS, Fahy E, Turnbull DM (2000). Transmission of the human mitochondrial genome. Hum. Reprod., 15(Suppl):235-245.

Iitsuka Y, Bock A, Nguyen DD, Samango-Sprouse CA, Simpson JL, Bischoff FZ (2001). Evidence of skewed X-chromosome inactivation in 47,XXY and 48,XXYY Klinefelter patients. Am J Med Genet. 98:25-31.

Imerslund O (1960). Idiopathic chronic megaloblastic anemia in children. Acta Paediatr. 49(Suppl):1-115.

International Human Genome Sequencing Consortium. Finishing the euchromatic sequence of the human genome (2004). Nature. 431:931-945.

Ireland H, Kunz G, Kyriakoulis K, Stubbs PJ, Lane DA (1997). Thrombomodulin gene mutations associated with myocardial infarction. Circulation. 96:15-18.

Ireland H, Konstantoulas CJ, Cooper JA, Hawe E, Humphries SE, Mather H, Goodall AH, Hogwood J, Juhan-Vague I, Yudkin JS, di Minno G, Margaglione M, Hamsten A, Miller GJ, Bauer KA, Kim YT, Stearns-Kurosawa DJ, Kurosawa S (2005). EPCR Ser219Gly: elevated sEPCR, prothrombin F1+2, risk for coronary heart disease, and increased sEPCR shedding in vitro. Atherosclerosis. 183:283-292.

Isermann B, Hendrickson SB, Hutley K, Wing M, Weiler H (2001). Tissue-restricted expression of thrombomodulin in the placenta rescues thrombomodulin-deficient mice from early lethality and reveals a secondary developmental block. Development. 128:827-838.

Jacks T (1996). Tumor suppressor gene mutations in mice. Annu Rev Genet. 30:603-36.

Jacobi FK, Leo-Kottler B, Mittelviefhaus K, Zrenner E, Meyer J, Pusch CM, Wissinger B (2001). Segregation patterns and heteroplasmy prevalence in Leber's hereditary optic neuropathy. Invest. Ophthalmol. Vis. Sci. 42:1208-1214.

Jenuth JP, Peterson AC, Fu K, Shoubridge EA (1996). Random genetic drift in the female germline explains the rapid segregation of mammalian mitochondrial DNA. Nat. Genet. 14:146-151.

Johnson TM, Hammond EM, Giaccia A, Attardi LD (2005). The p53QS transactivation-deficient mutant shows stress-specific apoptotic activity and induces embryonic lethality. Nat Genet. 37:145-152.

Jones SN, Roe AE, Donehower LA, Bradley A (1995). Rescue of embryonic lethality in Mdm2-deficient mice by absence of p53. Nature. 378:206-208.

Kalantry S, Manning S, Haub O, Tomihara-Newberger C, Lee HG, Fangman J, Disteche CM, Manova K, Lacy E (2001). The amnionless gene, essential for mouse gastrulation, encodes a visceral-endoderm-specific protein with an extracellular cysteine-rich domain. Nat Genet. 27:412-416.

Page 85: Genetic Studies on Recurrent Miscarriage

85

Kalousek DK, Pantzar T, Tsai M, Paradice B (1993). Early spontaneous abortion: morphologic and karyotypic findings in 3,912 cases. Birth Defects Orig Artic Ser. 29:53-61.

Katz VL, Kuller JA (1994). Recurrent miscarriage. Am J Perinat. 11:386-397.

Kay C, Jeyendran RS, Coulam CB (2006). p53 tumour suppressor gene polymorphism is associated with recurrent implantation failure. Reprod Biomed Online. 13:492-496.

Kere J (2001) Human population genetics: lessons from Finland. Annu Rev Genomics Hum Genet. 2:103-128.

Kesmodel U, Wisborg K, Olsen SF, Henriksen TB, Secher NJ (2002). Moderate alcohol intake in pregnancy and the risk of spontaneous abortion. Alcohol Alcohol. 37:87-92.

Khrapko K (2008). Two ways to make an mtDNA bottleneck. Nat. Genet. 40:134-135.

Kim JW, Park SY, Kim YM, Kim JM, Han JY, Ryu HM (2004). X-chromosome inactivation patterns in Korean women with idiopathic recurrent spontaneous abortion. J Korean Med Sci. 19:258-262.

Koehler CM, Lindberg GL, Brown DR, Beitz DC, Freeman AE, Mayfield JE, Myers AM (1991). Replacement of bovine mitochondrial DNA by a sequence variant within one generation. Genetics. 129:247-255.

Krausz C, McElreavey K (1999). Y chromosome and male infertility. Front Biosci. 15:E1-8.

Krawczak M, Reiss J, Cooper DN (1992). The mutational spectrum of single base-pair substitutions in mRNA splice junctions of human genes: causes and consequences. Hum Genet. 90:41-54.

Kujovich JL (2004). Thrombophilia and pregnancy complications. Am J Obstet Gynecol. 191:412-424.

Kuo PL (2002). Maternal trisomy 21 mosaicism and recurrent spontaneous abortion. Fertil Steril 78:432–433.

Kuo PL, Huang SC, Chang LW, Lin CH, Tsai WH, Teng YN (2008). Association of extremely skewed X-chromosome inactivation with Taiwanese women presenting with recurrent pregnancy loss. J Formos Med Assoc. 107:340-343.

Kupferminc M (2003). Thrombophilia and pregnancy. Reprod Biol Endocrinol 1:111.

Kutteh WH, Triplett DA (2006). Thrombophilias and recurrent pregnancy loss. Semin Reprod Med. 24:54-66.

Lacerda LL, Serrano SV, Mathes A, Rey JA, Bello MJ, Casartelli C (2005). An intronic variant in the TP53 gene in a Brazilian woman with breast cancer. Cancer Genet Cytogenet. 160:160-163.

Lanasa MC, Hogge WA, Kubik C, Blancato J, Hoffman EP (1999). Highly skewed X-chromosome inactivation is associated with idiopathic recurrent spontaneous abortion. Am J Hum Genet. 65:252-254.

Lanasa MC, Hogge WA, Kubik CJ, Ness RB, Harger J, Nagel T, Prosen T, Markovic N, Hoffman EP (2001). A novel X chromosome-linked genetic cause of recurrent spontaneous abortion. Am J Obstet Gynecol. 185:563–568.

Lander ES, Linton LM, Birren B, Nusbaum C, Zody MC, Baldwin J, Devon K, Dewar K, Doyle M, FitzHugh W, Funke R, Gage D, Harris K, Heaford A, Howland J, Kann L, Lehoczky J, LeVine R, McEwan P, McKernan K, Meldrim J, Mesirov JP, Miranda C, Morris W, Naylor J et al. (2001) Initial sequencing and analysis of the human genome. Nature. 409:860-921.

Larsson NG, Tulinius MH, Holme E, Oldfors A, Andersen O, Wahlstrom J, Aasly J (1992). Segregation and manifestations of the mtDNA tRNA(Lys) A-->G(8344) mutation of myoclonus epilepsy and ragged-red fibers (MERRF) syndrome. Am. J. Hum. Genet. 51:1201-1212.

Laskin CA, Bombsrdier C, Hannah ME, Mandel FP, Ritchie JWK, Farewell V, Farine D, Spitzer K, Fielding L, Soloninka C, Yeung M (1997). Prednisone and aspirin in women with autoantibodies and unexplained recurrent fetal loss. N Eng J Med. 337:148-153.

Page 86: Genetic Studies on Recurrent Miscarriage

86

Laszik Z, Mitro A, Taylor FB Jr, Ferrell G, Esmon CT (1997). Human protein C receptor is present primarily on endothelium of large blood vessels: implications for the control of the protein C pathway. Circulation. 96:3633-3640.

Lau AW, Brown CJ, Penaherrera M, Langlois S, Kalousek DK, Robinson WP (1997). Skewed X-chromosome inactivation is common in fetuses or newborns associated with confined placental mosaicism. Am J Hum Genet. 61:1353-1361.

Le Flem L, Picard V, Emmerich J, Gandrille S, Fiessinger JN, Aiach M, Alhenc-Gelas M (1999). Mutations in promoter region of thrombomodulin and venous thromboembolic disease. Arterioscler Thromb Vasc Biol. 19:1098-1104.

Lee KY, DeMayo FJ (2004). Animal models of implantation. Reproduction. 128:679-695.

Lehman TA, Haffty BG, Carbone CJ, Bishop LR, Gumbs AA, Krishnan S, Shields PG, Modali R, Turner BC (2000). Elevated frequency and functional activity of a specific germ-line p53 intron mutation in familial breast cancer. Cancer Res. 60:1062-1069.

Levine AJ (1997). p53, the cellular gatekeeper for growth and division. Cell. 88:323-331.

Levine JS, Branch DW, Rauch J (2002). The antiphospholipid syndrome. N Engl J Med. 346(10):752-763.

Levy S, Sutton G, Ng PC, Feuk L, Halpern AL, Walenz BP, Axelrod N, Huang J, Kirkness EF, Denisov G, Lin Y, MacDonald JR, Pang AW, Shago M, Stockwell TB, Tsiamouri A, Bafna V, Bansal V, Kravitz SA, Busam DA, Beeson KY, McIntosh TC, Remington KA, Abril JF, Gill J, Borman J, Rogers YH, Frazier ME, Scherer SW, Strausberg RL, Venter JC (2007). The diploid genome sequence of an individual human. PLoS Biol. 5:e254.

Li TC, Tuckerman EM, Laird SM (2002a). Endometrial factors in recurrent miscarriage. Hum Reprod Update. 8:43-52.

Li TC, Makris M, Tomsu M, Tuckerman EM, Laird SM (2002b). Recurrent miscarriage: aetiology, management and prognosis. Hum Reprod Update. 8:463-481.

Li W, Zheng X, Gu JM, Ferrell GL, Brady M, Esmon NL, Esmon CT (2005). Extraembryonic expression of EPCR is essential for embryonic viability. Blood. 106:2716-2722.

Lim CK, Cho JW, Song IO, Kang IS, Yoon YD, Jun JH (2008) Estimation of chromosomal imbalances in preimplantation embryos from preimplantation genetic diagnosis cycles of reciprocal translocations with or without acrocentric chromosomes. Fertil Steril. [Epub ahead of print]

Lim KS, Naviaux RK, Wong S, Haas RH (2008). Pitfalls in the denaturing high-performance liquid chromatography analysis of mitochondrial DNA mutation. J Mol Diagn. 10:102-108.

Livak KJ, Schmittgen TD (2001) Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods 25: 402-408.

Los FJ, Van Opstal D, van den Berg C (2004) The development of cytogenetically normal, abnormal and mosaic embryos: a theoretical model. Hum Reprod Update. 10:79-94.

Lucifero D, Chaillet JR, Trasler JM (2004). Potential significance of genomic imprinting defects for reproduction and assisted reproductive technology. Hum Reprod Update. 10:3-18.

Luder AS, Tanner SM, de la Chapelle A, Walter JH (2008). Amnionless (AMN) mutations in Imerslund-Gräsbeck syndrome may be associated with disturbed vitamin B(12) transport into the CNS. J Inherit Metab Dis. Jan 7. [Epub ahead of print]

Lyon MF (1961). Gene action in the X-chromosome of the mouse (Mus musculus L.). Nature. 190:372-373.

Lyon MF (1971). Possible mechanisms of X chromosome inactivation. Nat New Biol. 232:229–232.

Malcom S (2005). Genotype to Phenotype. Guide to mutation detection. Published by John Wiley &Sons, Inc.

Martinez-Barbera JP, Beddington RS (2001). Getting your head around Hex and Hesx1: forebrain formation in mouse. Int J Dev Biol. 45:327-336.

Page 87: Genetic Studies on Recurrent Miscarriage

87

Maruyama I, Bell CE, Majerus PW (1985). Thrombomodulin is found on endothelium of arteries, veins, capillaries, and lymphatics, and on syncytiotrophoblast of human placenta. J Cell Biol. 101:363-371.

Marzusch K, Ruck P, Horny HP, Dietl J, Kaiserling E (1995). Expression of the p53 tumour suppressor gene in human placenta: an immunohistochemical study. Placenta. 16:101-104.

McFadden DE, Friedman JM (1997) Chromosome abnormalities in human beings. Mutat Res. 396:129-140.

McFarland R, Clark KM, Morris AA, Taylor RW, Macphail S, Lightowlers RN, Turnbull DM (2002). Multiple neonatal deaths due to a homoplasmic mitochondrial DNA mutation. Nat. Genet. 30:145-146.

McKenzie M, Liolitsa D, Hanna MG (2004). Mitochondrial disease: mutations and mechanisms. Neurochem. Res. 29:589-600.

Medina P, Navarro S, Estelles A, Vaya A, Woodhams B, Mira Y, Villa P, Migaud-Fressart M, Ferrando F, Aznar J, Bertina RM, Espana F (2004). Contribution of polymorphisms in the endothelial protein C receptor gene to soluble endothelial protein C receptor and circulating activated protein C levels, and thrombotic risk. Thromb Haemost. 91:905-911.

Medina P, Navarro S, Estelles A, Vaya A, Bertina RM, Espana F (2005). Influence of the 4600A/G and 4678G/C polymorphisms in the endothelial protein C receptor (EPCR) gene on the risk of venous thromboembolism in carriers of factor V Leiden. Thromb Haemost. 94:389-394.

Mills JL, Simpson JL, Driscoll SG, Jovanovic-Petersen L, Van Allen M, Aarons JH, Metzger B, Bieber FR, Knopp RH, Holmes LB (1988). Incidence of spontaneous abortion among normal women and insulin dependent diabetic women whose pregnancies were identified within 21 days of conception. N Eng J Med. 319:1617-1623.

Minks J, Robinson WP, Brown CJ (2008) A skewed view of X chromosome inactivation. J Clin Invest. 118:20-23.

Montes de Oca Luna R, Wagner DS, Lozano G (1995). Rescue of early embryonic lethality in mdm2-deficient mice by deletion of p53. Nature. 378:203-206.

Moomjy M, Colombero LT, Veeck LL, Rosenwaks Z, Palermo GD (1999). Sperm integrity is critical for normal mitotic division and early embryonic development. Mol Hum Reprod. 5:836-844.

Mtiraoui N, Zammiti W, Ghazouani L, Braham NJ, Saidi S, Finan RR, Almawi WY, Mahjoub T (2006) Methylenetetrahydrofolate reductase C677T and A1298C polymorphism and changes in homocysteine concentrations in women with idiopathic recurrent pregnancy losses. Reproduction. 131:395-401.

Munne S, Alikani M, Tomkin G, Grifo J, Cohen J (1995). Embryo morphology, developmental rates, and maternal age correlated with chromosome abnormalities. Fertil Steril. 64:382-391.

Munne S, Escudero T, Sandalinas M, Sable D, Cohen J (2000). Gamete segregation in female carriers of Robertsonian translocations. Cytogenet Cell Genet. 90:303-308.

Mutter GL (1997). Role of imprinting in abnormal human development. Mutat Res. 396:141-147.

Nakai K, Sakamoto H (1994). Construction of a novel database containing aberrant splicing mutations of mammalian genes. Gene. 141:171-177.

Nakazawa F, Koyama T, Shibamiya A, Hirosawa S (2002). Characterization of thrombomodulin gene mutations of the 5'-regulatory region. Atherosclerosis. 164:385-387.

Naumova AK, Plenge RM, Bird LM, Leppert M, Morgan K, Willard HF, Sapienza C (1996). Heritability of X chromosome--inactivation phenotype in a large family. Am J Hum Genet. 58:1111-1119.

Naumova AK, Olien L, Bird LM, Smith M, Verner AE, Leppert M, Morgan K, Sapienza C (1998). Genetic mapping of X-linked loci involved in skewing of X chromosome inactivation in the human. Eur J Hum Genet. 6:552-562.

Page 88: Genetic Studies on Recurrent Miscarriage

88

Nicol CJ, Harrison ML, Laposa RR, Gimelshtein IL, Wells PG (1995). A teratologic suppressor role for p53 in benzo[a]pyrene-treated transgenic p53-deficient mice. Nat Genet. 10:181-187.

Nobrega MA, Pennacchio LA (2004). Comparative genomic analysis as a tool for biological discovery. J Physiol. 554:31-39.

Nolan PM, Hugill A, Cox RD (2002). ENU mutagenesis in the mouse: application to human genetic disease.Brief Funct Genomic Proteomic. 1:278-289.

Noordam MJ, van der Veen F, Repping S (2006). Techniques and reasons to remain interested in the Y chromosome. Fertil Steril. 86:1801-1802.

Noordam MJ, Repping S (2006). The human Y chromosome: a masculine chromosome. Curr Opin Genet Dev. 16:225-232.

Norimura T, Nomoto S, Katsuki M, Gondo Y, Kondo S (1996). p53-dependent apoptosis suppresses radiation-induced teratogenesis. Nat Med. 2:577-580.

Norio R (2003a). Finnish Disease Heritage I: characteristics, causes, background. Hum Genet. 112:441-456.

Norio R (2003b). Finnish Disease Heritage II: population prehistory and genetic roots of Finns. Hum Genet. 112:457-469.

Norwitz ER, Schust DJ, Fisher SJ (2001). Implantation and the survival of early pregnancy. N Engl J Med. 345:1400-1408.

Ntrivalas EI, Kwak-Kim JY, Gilman-Sachs A, Chung-Bang H, Ng SC, Beaman KD, Mantouvalos HP, Beer AE (2001). Status of peripheral blood natural killer cells in women with recurrent spontaneous abortions and infertility of unknown aetiology. Hum Reprod. 16(5):855-861.

Nybo Andersen AM, Wohlfahrt J, Christens P, Olsen J, Melbye M (2000). Maternal age and fetal loss: population based register linkage study. BMJ. 320:1708-1712.

Ober C, Aldrich CL, Chervoneva I, Billstrand C, Rahimov F, Gray HL, Hyslop T (2003). Variation in the HLA-G promoter region influences miscarriage rates. Am J Hum Genet. 72:1425-1435.

Ogasawara M, Aoki K, Okada S, Suzumori K (2000). Embryonic karyotype of abortuses in relation to the number of previous miscarriages. Fertil Steril. 73:300-304.

Okon MA, Laird SM, Tuckerman EM, Li TC (1998). Serum androgen levels in women who have recurrent miscarriages and their correlation with markers of endometrial function. Fertil Steril. 69:682-690.

Olivo PD, Van de Walle MJ, Laipis PJ, Hauswirth WW (1983). Nucleotide sequence evidence for rapid genotypic shifts in the bovine mitochondrial DNA D-loop. Nature. 306:400-402.

Otani T, Roche M, Mizuike M, Colls P, Escudero T, Munné S (2006) Preimplantation genetic diagnosis significantly improves the pregnancy outcome of translocation carriers with a history of recurrent miscarriage and unsuccessful pregnancies. Reprod Biomed Online. 13:869-874.

Pandey MK, Thakur S, Agrawal S (2004). Lymphocyte immunotherapy and its probable mechanism in the maintance of precnancy in women with recurrent spontaneous abortion. Arch Gynecol Obstet. 269:161-172.

Pandey MK, Rani R, Agrawal S (2005). An update in recurrent spontaneous abortion. Arch Gynecol Obstet. 272:95-108.

Parazzini F, Chatenoud L, Di Cintio E, Mezzopane R, Surace M, Zanconato G, Fedele L, Benzi G (1998). Coffee consumption and risk of hospitalized miscarriage before 12 weeks of gestation. Hum Reprod. 13:2286-2291.

Pegoraro E, Whitaker J, Mowery-Rushton P, Surti U, Lanasa M, Hoffman EP (1997). Familial skewed X inactivation: a molecular trait associated with high spontaneous-abortion rate maps to Xq28. Am J Hum Genet. 61:160-170.

Peltonen L, Jalanko A, Varilo T (1999) Molecular genetics of the Finnish disease heritage. Hum Mol Genet. 8:1913-1923.

Page 89: Genetic Studies on Recurrent Miscarriage

89

Peltonen L, McKusick VA (2001). Genomics and medicine. Dissecting human disease in the postgenomic era. Science. 291:1224-1229.

Percec I, Plenge RM, Nadeau JH, Bartolomei MS, Willard HF (2002). Autosomal dominant mutations affecting X inactivation choice in the mouse. Science. 296:1136-1139.

Perea-Gomez A, Vella FD, Shawlot W, Oulad-Abdelghani M, Chazaud C, Meno C, Pfister V, Chen L, Robertson E, Hamada H, Behringer RR, Ang SL (2002). Nodal antagonists in the anterior visceral endoderm prevent the formation of multiple primitive streaks. Dev Cell. 3:745-756.

Peters LL, Robledo RF, Bult CJ, Churchill GA, Paigen BJ, Svenson KL (2007). The mouse as a model for human biology: a resource guide for complex trait analysis. Nat Rev Genet. 8:58-69.

Pietrowski D, Bettendorf H, Riener EK, Keck C, Hefler LA, Huber JC, Tempfer C (2005). Recurrent pregnancy failure is associated with a polymorphism in the p53 tumour suppressor gene. Hum Reprod. 20:848-851.

Piette J, Neel H, Marechal V (1997). Mdm2: keeping p53 under control. Oncogene. 15:1001-1010.

Pim D, Banks L (2004). p53 polymorphic variants at codon 72 exert different effects on cell cycle progression. Int J Cancer. 108:196-199.

Plouffe L, White EW, Tho SP, Sweet CS, Layman LC, Whitman GF, McDonough PG (1992). Etiologic factors of recurrent abortion and subsequent reproductive performance of couples: Have we made any progress in the past 10 years? Am J Obstet Gynecol. 67:313-321.

Poland BJ, Miller JR, Jones DC, Trimble BK (1977). Reproductive counseling in patients who have had a spontaneous abortion. Am J Obstet Gynecol. 127:685-691.

Potdar N, Konje JC (2005).The endocrinological basis of recurrent miscarriages. Curr Opin Obstet Gynecol. 17:424-428.

Preston FE, Rosendaal FR, Walker ID, Breit E, Berntorp E, Conard J, Fontcuberta J, Makris M, Mariani G, Noteboom W, Pabinger I, Legnani C, Scharrer I, van der Meer FJ (1996). Increased fetal loss in women with heritable thrombophilia. Lancet. 348:913-916.

Prigoshin N, Tambutti M, Larriba J, Gogorza S, Testa R (2004). Cytokine gene polymorphisms in recurrent pregnancy loss of unknown cause. Am J Reprod Immunol. 52:36-41.

Puck JM, Willard HF (1998). X inactivation in females with X-linked disease. N Engl J Med. 338:325-328.

Quenby S, Vince G, Farquharson R, Aplin J (2002). Recurrent miscarriage: a defect in nature's quality control? Hum Reprod. 17:1959-1963.

Rai R, Regan L (2006). Recurrent miscarriage. Lancet. 368:601-611.

Rajasimha HK, Chinnery PF, Samuels DC (2008). Selection against pathogenic mtDNA mutations in a stem cell population leads to the loss of the 3243A-->G mutation in blood. Am J Hum Genet. 82:333-343.

Rand JH, Wu X-X, Andree HA, Lockwood CJ, Guller S, Scher J, Harpel PC (1997). Pregnancy loss in the antiphospholipid-antibody syndrome- a possible thrombogenic mechanism. N Eng J Med. 337(3):154-160.

Rebbeck TR, Spitz M, Wu X (2004). Assessing the function of genetic variants in candidate gene association studies. Nat Rev Genet. 5:589-597.

Regan L (1991). Recurrent miscarriage. BMJ. 302:543-544.

Regan L, Braude PR, Trembath PL (1989). Influence of past reproductive performance on risk of spontaneous abortion. BMJ. 299:541-545.

Reik W, Dean W (2001). DNA methylation and mammalian epigenetics. Electrophoresis. 22:2838-2843.

Reik W, Dean W, Walter J (2001). Epigenetic reprogramming in mammalian development. Science. 293:1089-1093.

Page 90: Genetic Studies on Recurrent Miscarriage

90

Reinhard G, Noll A, Schlebusch H, Mallmann P, Ruecker AV. Shifts in the TH1/TH2 balance during human pregnancy correlate with apoptotic changes (1998). Biochem Biophys Res Commun. 245(3):933-938.

Rey E, Kahn SR, David M, Shrier I (2003). Thrombophilic disorders and fetal loss: a meta-analysis. Lancet 361:901–908.

Reznikoff-Etiévan MF, Cayol V, Carbonne B, Robert A, Coulet F, Milliez J (2001) Factor V Leiden and G20210A prothrombin mutations are risk factors for very early recurrent miscarriage.BJOG. 108:1251-1254.

Risch HA, Weiss NS, Clarke EA, Miller AB (1988). Risk factors for spontaneous abortion and its recurrence. Am J Epidemiol. 128:420-430.

Roberts CJ, Lowe CR (1975). Where have all the conceptions gone? Lancet. i: 498-501.

Roberts CP, Murphy AA (2000). Endocrinopathies associated with recurrent pregnancy loss. Semin Reprod Med. 18:357–362.

Rogel A, Popliker M, Webb CG, Oren M (1985). p53 cellular tumor antigen: analysis of mRNA levels in normal adult tissues, embryos, and tumors. Mol Cell Biol. 5:2851-2855.

Rossant J, Cross JC (2001). Placental development: lessons from mouse mutants. Nat Rev Genet. 2:538-548.

Rubio C, Simon C, Blanco J, Vidal F, Minguez Y, Egozcue J, Crespo J, Remohi J, Pellicer A (1999). Implications of sperm chromosome abnormalities in recurrent miscarriage. J Assist Reprod Genet. 16:253-258.

Ruzzi L, Ciarafoni I, Silvestri L, Semeraro ML, Abeni D (2005). Association of PLA2 polymorphism of the ITGB3 gene with early fetal loss. Fertil Steril. 83:511-512.

Sadler TW, Langman J, Leland J (2006). Langman´s Medical Embryology. 10th edition, published

by Lippincott Williams and Wilkins.

Sah VP, Attardi LD, Mulligan GJ, Williams BO, Bronson RT, Jacks T (1995). A subset of p53-deficient embryos exhibit exencephaly. Nat Genet. 10:175-180.

Santos F, Dean W (2004). Epigenetic reprogramming during early development in mammals. Reproduction. 127:643-651.

Sariola H, Frilander M, Heino T, Jernvall J, Partanen J, Sainio K, Salminen M, Thesleff I (2003). 1

st edition, published by Gummerus Kirjapaino Oy.

Schmid P, Lorenz A, Hameister H, Montenarh M (1991). Expression of p53 during mouse embryogenesis. Development. 113:857-865.

Sharp A, Robinson D, Jacobs P (2000). Age- and tissue-specific variation of X chromosome inactivation ratios in normal women. Hum Genet. 107:343–349.

Shoubridge EA (2000). Mitochondrial DNA segregation in the developing embryo. Hum. Reprod. 15(Suppl):229-234.

Simerly C, Wu GJ, Zoran S, Ord T, Rawlins R, Jones J, Navara C, Gerrity M, Rinehart J, Binor Z (1995). The paternal inheritance of the centrosome, the cell's microtubule-organizing center, in humans, and the implications for infertility. Nat Med. 1:47-52.

Simmonds RE, Lane DA (1999). Structural and functional implications of the intron/exon organization of the human endothelial cell protein C/activated protein C receptor (EPCR) gene: comparison with the structure of CD1/major histocompatibility complex alpha1 and alpha2 domains. Blood. 94:632-641.

Simoni M, Bakker E, Krausz C (2004). EAA/EMQN best practice guidelines for molecular diagnosis of y-chromosomal microdeletions. Int J Androl. 27:240-249.

Snijders RJ, Sundberg K, Holzgreve W, Henry G, Nicolaides KH (1999). Maternal age- and gestation-specific risk for trisomy 21. Ultrasound Obstet Gynecol. 13:167-170.

Souza SS, Ferriani RA, Pontes AG, Zago MA, Franco RF (1999) Factor V leiden and factor II G20210A mutations in patients with recurrent abortion. Hum Reprod. 14:2448-1250.

Page 91: Genetic Studies on Recurrent Miscarriage

91

Stearns-Kurosawa DJ, Kurosawa S, Mollica JS, Ferrell GL, Esmon CT (1996). The endothelial cell protein C receptor augments protein C activation by the thrombin-thrombomodulin complex. Proc Natl Acad Sci U S A. 93:10212-10216.

Stephenson MD, Dreher K, Houlihan E, Wu W (1998). Prevention of unexplained recurrent spontaneous abortion using intravenous immunoglobulin: a prospective, randomized, double-blinded, placebo-controlled trial. Am J Reprod Immunol. 39:82-88.

Stephenson MD, Awartani KA, Robinson WP (2002). Cytogenetic analysis of miscarriages from couples with recurrent miscarriage: a case-control study. Hum Reprod. 17:446-451.

Stephenson M, Kutteh W (2007). Evaluation and management of recurrent early pregnancy loss. Clin Obstet Gynecol. 50:132-145.

Stern JJ, Dorfman AD, Gutierrez-Najar AJ, Cerrillo M, Coulam CB (1996). Frequency of abnormal karyotypes among abortuses from women with and without a history of recurrent spontaneous abortion. Fertil Steril. 65:250-253.

Stewart JB, Freyer C, Elson JL, Wredenberg A, Cansu Z, Trifunovic A, Larsson NG (2008). Strong purifying selection in transmission of mammalian mitochondrial DNA. PLoS Biol. 6:e10.

Stirrat GM (1990). Recurrent miscarriage. Lancet. 336:673-675.

Stogmann E, Lichtner P, Baumgartner C, Schmied M, Hotzy C, Asmus F, Leutmezer F, Bonelli S, Assem-Hilger E, Vass K, Hatala K, Strom TM, Meitinger T, Zimprich F, Zimprich A (2006). Mutations in the CLCN2 gene are a rare cause of idiopathic generalized epilepsy syndromes. Neurogenetics. 7:265-268.

Strachan T, Read AP (2004). Human Molecular Genetics. Third Edition, published by Garland Science.

Stray-Pedersen B, Stray-Pedersen S (1984). Etiologic factors and subsequent reproductive performance in 195 couples with a prior history of habitual abortion. Am J Obstet Gynecol. 148:140-146.

Strope S, Rivi R, Metzger T, Manova K, Lacy E (2004). Mouse amnionless, which is required for primitive streak assembly, mediates cell-surface localization and endocytic function of cubilin on visceral endoderm and kidney proximal tubules. Development. 131:4787-4795.

Suh Y, Vijg J (2005). SNP discovery in associating genetic variation with human disease phenotypes. Mutat Res. 573:41-53.

Sullivan AE, Lewis T, Stephenson M, Odem R, Schreiber J, Ober C, Branch DW (2003). Pregnancy outcome in recurrent miscarriage patients with skewed X chromosome inactivation. Obstet Gynecol. 101:1236-1242.

Sullivan AE, Silver RM, LaCoursiere DY, Porter TF, Branch DW (2004). Recurrent fetal aneuploidy and recurrent miscarriage. Obstet Gynecol. 104:784-788.

Suomalainen A, Syvänen AC (2000) Quantitative analysis of human DNA sequences by PCR and solid-phase minisequencing. Mol Biotechnol 15:123-131.

Suryanarayana V, Rao L, Kanakavalli M, Padmalatha V, Deenadayal M, Singh L (2006). Recurrent early pregnancy loss and endothelial nitric oxide synthase gene polymorphisms. Arch Gynecol Obstet. 274:119-124.

Swales AK, Spears N (2005). Genomic imprinting and reproduction. Reproduction. 130:389-399.

Takakuwa K, Adachi H, Hataya I, Ishii K, Tamura M, Tanaka K (2003). Molecular genetic studies of HLA-DRB1 alleles in patients with unexplained recurrent abortion in the Japanese population. Hum Reprod. 18:728-733.

Tal J, Kaplan M, Sharf M, Barnea ER (1991). Stress-related hormones affect human chorionic gonadothrophin secretion from the early human placenta in vitro. Hum Reprod. 6:766-769.

Tanner SM, Aminoff M, Wright FA, Liyanarachchi S, Kuronen M, Saarinen A, Massika O, Mandel H, Broch H, de la Chapelle A (2003). Amnionless, essential for mouse gastrulation, is mutated in recessive hereditary megaloblastic anemia. Nat Genet. 33:426-429.

Page 92: Genetic Studies on Recurrent Miscarriage

92

Tanner SM, Li Z, Bisson R, Acar C, Oner C, Oner R, Cetin M, Abdelaal MA, Ismail EA, Lissens W, Krahe R, Broch H, Grasbeck R, de la Chapelle A (2004). Genetically heterogeneous selective intestinal malabsorption of vitamin B12: founder effects, consanguinity, and high clinical awareness explain aggregations in Scandinavia and the Middle East. Hum Mut. 23:327-233.

Tay SK, Shanske S, Kaplan P, DiMauro S (2004). Association of mutations in SCO2, a cytochrome c oxidase assembly gene, with early fetal lethality. Arch Neurol. 61:950-952.

Taylor RW, Turnbull DM (2005). Mitochondrial DNA mutations in human disease. Nat. Rev. Genet. 6:389-402.

Teebi AS, Teebi SA (2005). Genetic diversity among the Arabs. Community Genet. 8:21-26.

Thomas M, Kalita A, Labrecque S, Pim D, Banks L, Matlashewski G (1999). Two polymorphic variants of wild-type p53 differ biochemically and biologically. Mol Cell Biol. 19:1092-1100.

Thorburn DR, Dahl HH (2001). Mitochondrial disorders: genetics, counseling, prenatal diagnosis and reproductive options. Am. J. Med. Genet. 106:102-114.

Tomihara-Newberger C, Haub O, Lee HG, Soares V, Manova K, Lacy E (1998). The amn gene product is required in extraembryonic tissues for the generation of middle primitive streak derivatives. Dev Biol. 204:34-54.

Tuckey RC (2005). Progesterone synthesis by the human placenta. Placenta. 26:273-281.

Tulppala M, Palosuo T, Ramsay T, Miettinen, Salonen R, Ylikorkala O (1993). A prospective study of 63 couples with history of recurrent spontaneous abortion: contributing factors and outcome of subsequent pregnancies. Hum Reprod. 8:764-770.

Uehara S, Hashiyada M, Sato K, Sato Y, Fujimori K, Okamura K (2001). Preferential X-chromosome inactivation in women with idiopathic recurrent pregnancy loss. Fertil Steril. 76:908-914.

Van Blerkom J (1996). Sperm centrosome dysfunction: a possible new class of male factor infertility in the human. Mol Hum Reprod. 2:349-354.

Van Blerkom J (2004). Mitochondria in human oogenesis and preimplantation embryogenesis: engines of metabolism, ionic regulation and developmental competence. Reproduction. 128:269-280.

Van Blerkom J, Sinclair J, Davis P (1998). Mitochondrial transfer between oocytes: potential applications of mitochondrial donation and the issue of heteroplasmy. Hum Reprod. 13:2857-2868.

Van Blerkom J, Davis P, Alexander S (2000). Differential mitochondrial distribution in human pronuclear embryos leads to disproportionate inheritance between blastomeres: relationship to microtubular organization, ATP content and competence. Hum Reprod. 15:2621-2633.

Van de Wouwer M, Collen D, Conway EM (2004). Thrombomodulin-protein C-EPCR system: integrated to regulate coagulation and inflammation. Arterioscler Thromb Vasc Biol. 24:1374-1383.

Van der Velden PA, Krommenhoek-Van Es T, Allaart CF, Bertina RM, Reitsma PH (1991). A frequent thrombomodulin amino acid dimorphism is not associated with thrombophilia. Thromb Haemost. 65:511-513.

Varley JM, Chapman P, McGown G, Thorncroft M, White GR, Greaves MJ, Scott D, Spreadborough A, Tricker KJ, Birch JM, Evans DG, Reddel R, Camplejohn RS, Burn J, Boyle JM (1998). Genetic and functional studies of a germline TP53 splicing mutation in a Li-Fraumeni-like family. Oncogene. 16:3291-3298.

Venter JC, Adams MD, Myers EW, Li PW, Mural RJ, Sutton GG, Smith HO, Yandell M, Evans CA, Holt RA, Gocayne JD, Amanatides P, Ballew RM, Huson DH, Wortman JR, Zhang Q, Kodira CD, Zheng XH, Chen L, Skupski M, Subramanian G, Thomas PD, Zhang J et al. (2001) The sequence of the human genome. Science. 291:1304-1351.

Vitiello D, Patrizio P (2007). Implantation and early embryonic development: implications for pregnancy. Semin Perinatol. 31:204-207.

Page 93: Genetic Studies on Recurrent Miscarriage

93

Voglino G, Castello S, Silengo L, Stefanuto G, Friard O, Ferrara G, Fessia L (1997). An intronic deletion in TP53 gene causes exon 6 skipping in breast cancer. Eur J Cancer. 33:1479-1483.

Vogt PH, Edelmann A, Kirsch S, Henegariu O, Hirschmann P, Kiesewetter F, Kohn FM, Schill WB, Farah S, Ramos C, Hartmann M, Hartschuh W, Meschede D, Behre HM, Castel A, Nieschlag E, Weidner W, Grone HJ, Jung A, Engel W, Haidl G (1996). Human Y chromosome azoospermia factors (AZF) mapped to different subregions in Yq11. Hum Mol Genet. 5:933-943.

Von Linsingen R, Bompeixe EP, Bicalho G (2005). A case-control study in IL6 and TGFB1 gene polymorphisms and recurrent spontaneous abortion in southern Brazilian patients. Am J Reprod Immunol. 53:94-99.

Wan J, Carr JR, Baloh RW, Jen JC (2005). Nonconsensus intronic mutations cause episodic ataxia. Ann Neurol. 57:131-135.

Wang GJ, Yang P, Xie HG (2006). Gene variants in noncoding regions and their possible consequences. Pharmacogenomics. 7:203-209.

Wang H, Dey SK (2006). Roadmap to embryo implantation: clues from mouse models. Nat Rev Genet. 7:185-199.

Wang JX, Davies MJ, Norman RJ (2002). Obesity increases the risk of spontaneous abortion during infertility treatment. Obes Res. 10:551-554.

Wang X, Bornslaeger EA, Haub O, Tomihara-Newberger C, Lonberg N, Dinulos MB, Disteche CM, Copeland N, Gilbert DJ, Jenkins NA, Lacy E, (1996). A candidate gene for the amnionless gastrulation stage mouse mutation encodes a TRAF-related protein. Dev Biol. 177:274-290.

Wang X, Lin Q, Ma Z, Hong Y, Zhao A, Di W, Lu P (2005). Association of the A/G polymorphism at position 49 in exon 1 of CTLA-4 with the susceptibility to unexplained recurrent spontaneous abortion in the Chinese population. Am J Reprod Immunol. 53:100-105.

Wang XB, Zhao X, Giscombe R, Lefvert AK (2002). A CTLA-4 gene polymorphism at position -318 in the promoter region affects the expression of protein. Genes Immun. 3:233-234.

Wang-Gohrke S, Weikel W, Risch H, Vesprini D, Abrahamson J, Lerman C, Godwin A, Moslehi R, Olipade O, Brunet JS, Stickeler E, Kieback DG, Kreienberg R, Weber B, Narod SA, Runnebaum IB (1999). Intron variants of the p53 gene are associated with increased risk for ovarian cancer but not in carriers of BRCA1 or BRCA2 germline mutations. Br J Cancer. 81:179-183.

Warren JE, Silver RM (2008). Genetics of pregnancy loss. Clin Obstet Gynecol. 51:84-95.

Waterston RH, Lindblad-Toh K, Birney E, Rogers J, Abril JF, Agarwal P, Agarwala R, Ainscough R, Alexandersson M, An P, Antonarakis SE, Attwood J, Baertsch R, Bailey J, Barlow K, Beck S, Berry E, Birren B, Bloom T, Bork P, Botcherby M, Bray N, Brent MR, Brown DG et al. (2002). Initial sequencing and comparative analysis of the mouse genome. Nature. 420:520-562.

Wegmann TG, Lin H, Guilbert L, Mosmann TR (1993). Bidirectional cytokine interactions in the maternal-fetal relationship: is successful pregnancy a TH2 phenomenon? Immunol Today. 14(7):353-356.

Weiler-Guettler H, Aird WC, Rayburn H, Husain M, Rosenberg RD (1996). Developmentally regulated gene expression of thrombomodulin in postimplantation mouse embryos. Development. 122:2271-2281.

Weinstein LB (2007). Selected genetic disorders affecting Ashkenazi Jewish families. Fam Community Health. 30:50-62.

Wilcox AJ, Weinberg CR, O'Connor JF, Baird DD, Schlatterer JP, Canfield RE, Armstrong EG, Nisula BC (1988). Incidence of early loss of pregnancy. N Engl J Med. 319:189-194.

Wong LJ (2007). Pathogenic mitochondrial DNA mutations in protein-coding genes. Muscle Nerve. 36:279-293.

Page 94: Genetic Studies on Recurrent Miscarriage

94

Xiao W, Oefner PJ (2001). Denaturing high-performance liquid chromatography. Hum Mutat. 17:439-474.

Yamada H, Kato EH, Kobashi G, Ebina Y, Shimada S, Morikawa M, Sakuragi N, Fujimoto S (2001). High NK cell activity in early pregnancy correlates with subsequent abortion with normal chromosomes in women with recurrent spontaneous abortion. Am J Reprod Immunol. 46(2):132-136.

Yang Y, Ma M, Li L, Zhang W, Xiao C, Li S, Ma Y, Tao D, Liu Y, Lin L, Zhang S (2008). Evidence for the association of Y-chromosome haplogroups with susceptibility to spermatogenic failure in a Chinese Han population. Med Genet. 45:210-215.

Zhu M, Zhao S (2007) Candidate gene identification approach: progress and challenges. Int J Biol Sci. 3:420-427.